Lecture Notes On Linear Algebra: Prepared by Muhammad Shahnewaz Bhuyan
Lecture Notes On Linear Algebra: Prepared by Muhammad Shahnewaz Bhuyan
Prepared by
Muhammad Shahnewaz Bhuyan
Department of Mathematics
Chittagong University of Engineering and Technology
Chattogram-4349
Syllabus
Matrices and elementary row operations, Rank, Inverse of a square matrix.
System of linear equations and their applications and their applications in network flow and
electric circuits.
Vectors in Rn , Linear combinations, Linear dependence and independence
Eigenvalues and eigenvectors, Diagonalization.
Books Recommended
(i) H. Anton and C. Rorres, Elementary Linear Algebra (Applications Version), Jhon Wiley &
Sons, Eleventh Edition, 2020-21.
(ii) S. Lipschutz and M. L. Lipson, Theory and Problems of Linear Algebra, Schaum’s Outline
Series, McGraw-Hill, Third Edition, 2013-2014.
(iii) M. Abdur Rahman, College Linear Algebra: Theory of Matrices with Applications, Nahar
Book Depot & Publications, Dhaka, Sixth Edition (Reprint), 2012.
Class Routine
Monday
11:00 AM — 11:50 AM (Section B)
11:50 AM — 12:40 PM (Section A)
Tuesday
11:00 AM — 11:50 AM (Section B)
11:50 AM — 12:40 PM (Section A)
ii
Contents
Outline ii
I Linear Algebra 1
1 Matrices 2
1.1 Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.1.1 Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.1.2 Order of a matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.1.3 Equality of two matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.2 Matrix Operations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2.1 Sum of two matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2.2 Subtraction of two matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2.3 Scalar multiplication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
1.2.4 Matrix multiplication . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.2.5 Visualizing the effect of matrix multiplication on plane . . . . . . . . . . . . 6
1.3 Differences between Real Arithmetic and Matrix Arithmetic . . . . . . . . . . . . . 6
1.3.1 Failure of multiplicative commutativity . . . . . . . . . . . . . . . . . . . . . 6
1.3.2 Failure of cancellation Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.3.3 Matrix arithmetic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.4 Some Special Types of Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.4.1 Row matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.4.2 Column matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.4.3 Zero matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.4.4 Zero product with nonzero factors . . . . . . . . . . . . . . . . . . . . . . . . 9
1.4.5 Identity matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.4.6 Diagonal matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.4.7 Lower triangular matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.4.8 Upper triangular matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.4.9 Scalar matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.4.10 Transpose matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.4.11 Symmetric matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.4.12 Skew-symmetric matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.4.13 Orthogonal matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.4.14 Involutory matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.4.15 Idempotent matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.4.16 Nilpotent matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.4.17 Periodic matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.4.18 The Vandermonde matrix of order m . . . . . . . . . . . . . . . . . . . . . . 14
1.5 Square Root of a Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
iii
1.6 Matrix Polynomial . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.7 Application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.7.1 A special set . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2 Determinants 16
2.1 Determinants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.1.1 Determinants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.2 Evaluating Determinants by Cofactor Expansion . . . . . . . . . . . . . . . . . . . . 16
2.2.1 Determinant of a 1 × 1 matrix . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.2.2 Determinant of a 2 × 2 matrix . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.2.3 Minors and cofactors of a square matrix . . . . . . . . . . . . . . . . . . . . 17
2.2.4 Determinant of a general n × n square matrix . . . . . . . . . . . . . . . . . 19
2.3 Determinants for Solving a System of Linear Equations . . . . . . . . . . . . . . . . 21
2.4 Some Properties of Determinants . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.4.1 Some Properties of Determinants . . . . . . . . . . . . . . . . . . . . . . . . 23
2.5 Construction of a Square Matrix with Given Order and Determinant . . . . . . . . 28
2.5.1 Construction of a n × n matrix with determinant ∆ . . . . . . . . . . . . . . 28
2.6 Inverse of a Square Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.6.1 Cofactor matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.6.2 Adjoint matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.6.3 Another approach for finding inverse matrix . . . . . . . . . . . . . . . . . . 33
iv
5 Vector Spaces 44
5.1 Vectors in Rn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
5.1.1 Lines and planes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
5.1.2 Norm, dot product and distance . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.1.3 Distance between two vectors . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.1.4 Normalizing a vector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.1.5 Euclidean inner product . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
5.1.6 Inequalities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.2 Vector Spaces and Subspaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
5.2.1 Real vector space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.2.2 Vector subspaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.3 Linear Independence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.3.1 Linear combination of vectors . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.3.2 Linear independence of vectors . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.4 Basis and Dimension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.4.1 Spanning set and basis of a vector space . . . . . . . . . . . . . . . . . . . . 47
5.4.2 Dimension . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
5.5 Row Spaces and Column Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
5.6 Rank and Nullity of a Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
8 Linear Transformation 57
8.1 Linear Transformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
8.1.1 Reflection about x-axis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
8.1.2 Reflection about line y = x . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
8.1.3 Combination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
8.2 Kernel and Range . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
8.3 Rank and Nullity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
8.4 Isomorphism of Vector Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
8.4.1 One-one . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
8.4.2 Onto . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
8.4.3 Inverse transformation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
8.4.4 Isomorphism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
8.5 Matrices for General Linear Transformation . . . . . . . . . . . . . . . . . . . . . . 58
v
9 Quadratic Forms 60
9.1 Quadratic Forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
9.2 Application . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
Bibliography 69
vi
Part I
Linear Algebra
1
Chapter 1
Matrices
For detailed explanations and solving more problems, readers are referred to Anton and Ror-
res [2], Hamilton [4], Hadley [3], Larson and Falvo [5], Lipschutz and Lipson [6], Abdur Rahman[7],
Wildgerger [8].
1.1 Matrices
1.1.1 Matrices
A matrix is nothing but a rectangular array consisting of some horizontal and some vertical lines of
numbers. Each number of that array is referred as an entry of that matrix, when horizontal lines of
numbers are referred as rows and vertical lines of numbers are referred as columns of that matrix.
Definition 1.1.1 A matrix A with m rows and n columns is a rectangular array of numbers,
represented in either of the forms
a11 a12 a13 · · · a1n a11 a12 a13 ··· a1n
a21 a22 a23 · · · a2n a21 a22 a23 ··· a2n
A = a31 a32 a33 · · · a3n or A = a31 a32 a33 ··· a3n
.
.. .. .. .. .. .. .. .. .. ..
. . . . . . . . . .
am1 am2 am3 · · · amn am1 am2 am3 ··· amn
Here
a11
a12 a13 ··· a1n is called the first row of A
a21
a22 a23 ··· a2n is called the second row of A
a31
a32 a33 ··· a3n is called the third row of A
2
Remark 1.1.1 In this text, we use the symbol for matrix notation.
Definition 1.1.2 In a matrix A = aij , each aij is called an entry located at the intersection
of i-th row and j-th column. The entry aij is also referred as (i, j)-th entry.
0 1 4
Example 1.1.1 The rectangular array of real numbers is a matrix.
3 13 5
Here A is not a matrix, as it’s 2nd row 2nd column position is empty. Again, B is not a matrix, as
in it’s 1st row 3rd column position there is ∞ symbol, which is not a number.
Definition 1.1.3 Matrix A = aij is said to be a real matrix, if all aij ∈ R.
1 4 √ 1 4
Example 1.1.2 A = is a real matrix, but B = is not a real matrix,
−4 8 −4 8
√
as −4 is an entry of B, which does not belong to R.
Remark 1.1.2 In this text, unless otherwise specified explicitly, all matrices are assumed to be
real.
Question 1.1.1 Is a213 the (21, 3)-th entry or (2, 13)-th entry in aij ?
Answer It depends on the previous entry’s row number and column number.
Definition 1.1.4 A matrix having equal number of rows and columns is called a square matrix.
Definition 1.1.5 Let A = aij be a square matrix of order n × n. Then the entries aij for which
i = j form the principal diagonal of A and are called the diagonal entries of A.
2 3 5
Example 1.1.3 1 0 6 is a square matrix in which 2, 0 and −2 are the diagonal
0 0 −2
2 3 5
entries. On the other hand is not a square matrix.
1 0 6
3
Definition 1.1.6 Let A = aij and B = bij be two matrices. Then A and B are said to be
equal, if for each i = 1, 2, 3, . . . , m and j = 1, 2, 3, . . . , n,
aij = bij .
To denote two matrices A and B are equal we write A = B whereas to denote two matrices A and
B are not equal, we write A ̸= B.
Here A and D are equal as their orders and corresponding entries are same. Again, A and B are
not equal though they have same order, as (1, 1)-th entry of A is 1 whereas (1, 1)-th entry of B is
5. Moreover, A and C are not equal, are their orders are different.
On the other hand, A + C is not defined, as their orders are not same.
On the other hand, A − C is not defined, as their orders are not same.
4
1 3 4
Example 1.2.3 Suppose that A = 5
6 7 and α = 4. Then
3 3 4
1 3 4 4 12 16
αA = 4 5 0 7 = 20 0 28 .
0 3 −4 0 12 −16
l
X
cij = aik bkj
k=1
= ai1 b1j + ai2 b2j + ai3 b3j + · · · + ain bnj , 1 ≤ i ≤ m, 1 ≤ j ≤ n .
The order of C = AB is m × n.
Formula used in the above definition is known as row-column rule for matrix multiplication.
Note 1.2.1 With the help of the above row-column formula for matrix multiplication, it is possible
to find a specific entry of a product matrix without determining the product matrix fully.
1 3 4 3 3 4
Example 1.2.4 Suppose that A = and B = . Here AB is not
5 6 0 5 0 7
defined, as the number of columns of A is not equal to the number of rows of B.
3 3
1 3 4
Example 1.2.5 Suppose that A = and B = −1 0 . Here AB is defined,
5 6 0
1 0
as the number of columns of A is equal to the number of rows of B. Say AB = C. Clearly, C is a
1
The notion of matrix multiplication was first introduced by German mathematician Gotthold Eisenstein, a
student of Gauss, around 1844. And this concept was expanded and formalized further by Cayley in 1858. Eisentstien
was suffering from bad health throughout his life and died when he was only 30. So his potential was never realized
[2, Page 30].
5
matrix of order 2 × 2. Now,
3
X
c11 = a1k bk1 = a11 b11 + a12 b21 + a13 b31 = 1(3) + 3(−1) + 4(1) = 4,
k=1
X3
c12 = a1k bk2 = a11 b12 + a12 b22 + a13 b32 = 1(3) + 3(0) + 4(0) = 3,
k=1
X3
c21 = a2k bk1 = a21 b11 + a22 b21 + a23 b31 = 5(3) + 6(−1) + 0(1) = 5,
k=1
3
X
c22 = a2k bk2 = a21 b12 + a22 b22 + a23 b32 = 5(3) + 6(0) + 0(0) = 15.
k=1
4 3
Therefore C = .
9 15
3 3
1 3 0
Problem 1.2.1 Find the product of A = and B = −1 0 .
5 6 0
1 0
Solution Here
3 3
1 3 0
AB = . −1 0
5 6 0
1 0
1(3) + 3(−1) + 0(1) 1(3) + 3(0) + 0(0) 0 3
= = .
5(3) + 6(−1) + 0(1) 5(3) + 6(0) + 0(0) 9 15
Problem 1.3.1 Find two such matrices A and B for which AB ̸= BA.
6
Proof. We have A2 = A.A. But A.A is defined only when A is a square matrix.
Problem 1.3.2 Prove or disprove that (AB)2 = A2 B 2 for any two square matrices A and B.
2 1 1 3
Disproof Suppose that A = and B = . Here
3 2 4 1
2 85 119 2 2 123 94
(AB) = , whereas A B = .
187 198 250 241
Theorem 1.3.1 Let α, β be two scalars and A, B, C be three matrices of suitable order so that
indicated operations are defined. Then
(f) α(A + B) = αA + αB
(g) (α + β)A = αA + βA
7
Question 1.3.2 If A2 + AB − 2(A + B) − I3 is defined, then what are the order of it and A?
Question 1.4.1 Is there any matrix which is at the same time a row matrix and a column matrix?
Answer Infinitely many. For example, 4 , −4 , 0 etc.
Theorem 1.4.1 Let α ∈ R be a scalar and 0 be a zero matrix. Then for a matrix A with suitable
order for which corresponding matrix operation to be defined
(a) A + 0 = 0 + A = A.
(b) A − A = A + (−A) = 0.
(c) A − 0 = A.
(d) A0 = 0A = 0.
8
1.4.4 Zero product with nonzero factors
In case of usual algebra of real numbers, product of two nonzero numbers can not be zero. But
in algebra of matrices, product of two nonzero matrices can be a zero matrix. Following example
demonstrates this fact.
0 −1 10 1
Example 1.4.4 For the matrices A = and B = , though AB = 0, neither
0 3 0 0
A = 0 nor B = 0.
Question 1.4.2 For two matrices A and B, does AB = 0 imply either A = 0 or B = 0? Justify
your answer.
Counterexample 1.4.2 Consider the matrices A and B of Example 1.4.4. Here though AB = 0,
neither A = 0 nor B = 0.
9
4 0 0
Example 1.4.8 1 0 0 is a lower triangular matrix.
0 1 −1
That is, in an upper triangular matrix all entries below the diagonal are zero.
4 0 5
Example 1.4.9 0 0 1 is a upper triangular matrix.
0 0 −1
Note 1.4.2 Zero matrix and identity matrix are at a time upper triangular and lower triangular.
AT = aji .
Theorem 1.4.3 Let A and B be two matrices of suitable orders to define the matrix operations
between them. Then
T
(a) AT = A.
(b) (A + B)T = AT + B T .
(c) (AB)T = B T AT .
(d) (αA)T = αAT for a scalar α.
Question 1.4.3 (Should be rectified) Can the property (AB)T = B T AT be extended for 3 matri-
ces? Can this property be extended to finite number of matrices? What is about infinite number
of matrices?
10
Answer Yes. Yes. Because product of any two matrices is again a matrix.
In case of infinite number of matrices it can not be said that from where it will start back.
is a symmetric matrix, as
4 5 6 0
5 3 0 10
AT = = A.
6 0 7 −1
0 10 −1 8
AT = −A.
Obviously, every skew-symmetric matrix is a square matrix. In a skew-symmetric matrix all entries
along its principal diagonal are 0 (zero).
is a skew-symmetric matrix, as
0 −5 −6 0
5 0 0 −10
AT =
6
= −A.
0 0 1
0 10 −1 0
11
is a skew-symmetric matrix, as
0 −5 −6 8
5 0 4 −10
AT = = −A.
6 −4 0 1
−8 10 −1 0
AAT = AT A = I.
A−1 = AT .
Obviously, every orthogonal matrix is a square matrix. All identity matrices are orthogonal.
Note 1.4.4 Collection O of all orthogonal matrices of order m × m forms a group, called or-
thogonal group.
12
1.4.15 Idempotent matrix
Let A be a square matrix. Then A is said to be idempotent, if A2 = A.
1
1 1
−1 3 5 2
Example 1.4.16 The matrices 1 −3 −5, 21 2
1 , 1
2
1 etc are idempo-
−1 3 5
8 2 2 2
tent.
2 −2 −4
Problem 1.4.2 Show that −1 3 4 is an idempotent matrix.
1 −2 −3
1
Example 1.4.17 Any n × n matrix with the entries of the form is idempotent.
n
Problem 1.4.3 If A is a idempotent matrix, then what does equal to An for n ∈ N?
Solution A.
Problem 1.4.4 Write down a 4 × 4 nonzero idempotent matrix A for which |A| = 0.
1 1 1 1
4 4 4 4
1 1 1 1
4 4 4 4
Solution The matrix 1 is of such form.
1 1 1
4 4 4 4
1 1 1 1
4 4 4 4
Proposition 1.4.4 If A is an idempotent matrix, then I − A is also idempotent.
(I − A)2 = (I − A)(I − A)
= I 2 − IA − AI + A2
=I −A−A+A
= I − A.
Theorem 1.4.4 Let A and B be two idempotent matrices. Then AB is idempotent, when
AB = BA.
A2 = A and B 2 = B.
Here
Hence AB is idempotent.
13
Theorem 1.4.5 Let A and B be two idempotent matrices. Then A + B is idempotent if and
only if AB = BA = 0.
Theorem 1.4.6 Let A and B be two square matrices order n with AB = A and BA = B. Then
A and B are als0 idempotent.
An = 0.
An+1 = A.
14
Problem 1.5.1
[2,Like Problem 29 of Exerecise Set 1.3, Page 38] Find the square root of the
1 1
matrix B = .
0 1
1! 1!
1 1 1 1
Solution A2 = 2 2 = 0 .
0 1 0 1 1
Problem 1.5.2 Problem 29(a) and Problem 29(b) [2, Exerecise Set 1.3, Page 38]
1.7 Application
Problem 1.7.1 Suppose that three students Jone, Pitar and Strang obtained marked in three
assignments A20, A10, A10 and in class performance C05 as the following Table 1.1.
Assignments
Student’s Name A20 A10 A10 C05
Jone 20 07 9.5 03
Pitar 19 08 09 05
Strang 10 06 07 00
Tk per number 05 03 04 10
Table 1.1:
If anyone get Tk 05, Tk 03, Tk 03 and Tk 04 for each mark in A20, A10, A10 and C05
respectively, then calculate the amount of money received by Jone, Pitar and Strang.
15
Chapter 2
Determinants
For detailed explanations and solving more problems, readers are referred to Anton and Rorres [2],
Hamilton [4], Hadley [3], Larson and Falvo [5], Lipschutz and Lipson [6].
2.1 Determinants
2.1.1 Determinants
Determinant is defined for square matrices. Corresponding to every square matrix there exists a
unique number, which is referred as it’s determinant.
Determinant of a square matrix may be negative, positive or zero. Every square matrix relates
to a unique determinant. That is why, determinants can also be interpreted as a function.
Repelling a misconception
Sometimes it is said that determinant is the value of a square matrix. This is a misconception.
Determinant is not the value of a matrix, rather it is simply a number related to a square matrix.
16
Example 2.2.1 Consider the 1 × 1 matrix A = −2 . Then
−2 = −2.
−2 0
= (−2)(3) − (2)(0) = −6 .
2 3
Remark 2.2.1 From here of this text, we use the notation det(A) only with the name A of the
square matrix A = aij and the notation aij elsewhere.
M12 = 8 = 8,
M21 = −4 = −4
M22 = 1 = 1.
17
1 −4 4
Example 2.2.4 Consider the matrix A = 5 0 8 . Here minor corresponding to the
1
−5 −9 2
(1, 1)-th entry 1 is
0 8
M11 = 1 = 0 − (−72) = 72,
−9 2
minor corresponding to the (1, 2)-th entry −4 is
5 8 5 85
M12 = 1 = − (−40) = ,
−5 2 2 2
minor corresponding to the (1, 3)-th entry 4 is
5 0
M12 = = (−45) − (0) = −45,
−5 −9
and similarly
M21 = −38, M22 = 392
, M23 = 11,
M31 = −32, M32 = −12, M33 = 20.
Definition 2.2.4 Let A = aij be square matrix and Mij be the corresponding minor of the
entry aij . Then the cofactor corresponding to the entry aij is denoted by Cij and defined as
Clearly, cofactor corresponding to an entry of a square matrix is nothing but the corresponding
signed minor and so cofactor is also a number.
Example
2.2.5 In Example
2.2.4 we obtained the minors of corresponding entries of the matrix
1 −4 4
A= 5 0 8 as
1
−5 −9 2
and similarly
C21 = 38, C22 = 39
2
, C23 = −11,
C31 = −32, C32 = 12, C33 = 20.
Here observe that the cofactor corresponding to any entry of the matrix A, is either same or opposite
to its corresponding minor.
18
Note 2.2.1 The sign of the cofactor corresponding to an entry of a square matrix is either same
or opposite to the sign of its corresponding minor.
displays the rule (−1)i+j for assigning sign of the cofactor Cij corresponding to the entry aij of a
square matrix. According to the pattern of this checkerboard array,
and so on.
Observe that in the checkerboard array along the principal diagonal all signs are positive.
Note 2.2.3 The sum of the products of the elements of a row (or a column) and the respective
10 5
cofactors of any other row (or any other column) is zero. For example, in A = we
3 −8
consider the elements of 1st row and the corresponding cofactors of 2nd row to obtain that
or
det(A) = a1j C1j + a2j C2j + a3j C3j + · · · + anj Cnj .
| {z }
cofactor expansion along the j−th column
−2 0
Example 2.2.6 Consider the matrix A = given in Example 2.2.2 . All minors and
2 3
the cofactors of A are respectively
and
C11 = 3, C12 = −2, C21 = −0 = 0, C22 = −2,
Taking cofactor expansion along the 1st row of A, we obtain
19
which also coincides with the det(A) obtained in Example 2.2.2 .
Similarly, we get the same value of det(A), taking cofactor expansion along the 2nd row and
2nd column of A.
−2 0 1
Example 2.2.7 Consider the matrix A = 2 3 0 . Here cofactor corresponding to
−1 −3 4
the (1, 1)-th entry −2 of A is
3 0
C11 = (−1)1+1 · = (1)(12 − 0) = 12,
−3 4
cofactor corresponding to the (1, 2)-th entry 0 of A is
2 0
C12 = (−1)1+2 · = (−1)(8 − 0) = −8,
−1 4
cofactor corresponding to the (1, 3)-th entry 1 of A is
2 3
C13 = (−1)1+3 · = (1) ((−6) − (−3)) = −3,
−1 −3
and similarly
C21 = −3, C22 = −7, C23 = −6,
C31 = −3, C32 = 2, C33 = −6.
Taking cofactor expansion along the 1st row of A, we obtain
det(A) = a11 C11 + a12 C12 + a13 C13
= (−2)(12) + (0)(−8) + (1)(−3) = −27.
Again, taking cofactor expansion along the 1st column of A, we obtain
det(A) = a11 C11 + a21 C21 + a31 C31
= (−2)(12) + (2)(−3) + (−1)(−3) = −27,
which coincides with the det(A) obtained by taking cofactor expansion along the 1st row of A.
Similarly, we get the same value of det(A), taking cofactor expansion any row and any column
of A.
0 −2 1
Problem 2.2.1 For the matrix A = −1 3 0 ,
2 −3 3
a. find det(A), taking cofactor expansion along the 1st row.
b. find det(A), taking cofactor expansion along the 2nd column.
c. compare the values of det(A) obtained by cofactor expansion along the 1st row and 2nd column
of A.
20
b. Taking cofactor expansion along the 2nd column of A, we obtain
c. Values of det(A) obtained by cofactor expansion along the 1st row and 2nd column of A
are equal.
0 70 1000
Exercise 2.2.1 To find the determinant of the matrix A = 5 0 10 , cofactor ex-
8 0 2
pansion along which row or column will be comparatively easy? Why? Find det(A) also.
determinant of two different matrices with different order never become equal.
or, determinant of two matrices are equal implies these are of same order.
a1 x + b 1 y = c 1 (2.1)
a2 x + b 2 y = c 2 (2.2)
is
Dx Dy
x= , y= ,
D D
where
a1 b1 c1 b1 a1 c1
D= , Dx = , Dy = .
a2 b2 c2 b2 a2 c2
a1 b 2 x + b 1 b 2 y = b 2 c 1 (2.3)
a2 b 1 x + b 1 b 2 y = b 1 c 2 (2.4)
1
Though variations of Cramer’s rule were fairly well know before it was discussed in the research work of the
Swiss mathematician Gabriel Cramer (1704-1752) in the year 1750, it was popularised by him. So mathematicians
attached his name with this method (see [2, Page 124]).
21
Subtracting Equation (2.4) from Equation (2.3), we find
b2 c 1 − b1 c 2
x=
a1 b 2 − a2 b 1
c1 b1
c2 b2 Dx
= = .
a1 b1 D
a2 b2
a1 a2 x + a2 b1 y = a2 c1 (2.5)
a1 a2 x + a1 b2 y = a1 c2 (2.6)
Problem 2.3.1 Using Cramer’s Rule, solve the following system of linear equations
2x + 3y = 1
3x − 4y = 3
2 3 1 3 2 1
D= = −17, Dx = = −13, Dy = = 3.
3 −4 3 −4 3 3
Thus
Dx 13 Dy 3
x= = , y= =− .
D 17 D 17
Therefore our required solution is
13 3
(x, y) = ,− .
17 17
13 3
Answer. (x, y) = 17
, − 17 .
22
Theorem 2.3.2 (Cramer’s rule for solving a system of linear equations in three vari-
ables) Solution, if exists, of the system of linear equations in three variables
a1 x + b1 y + c1 z = d1 (2.8)
a2 x + b2 y + c2 z = d2 (2.9)
a3 x + b 3 y + c 3 z = d 3 (2.10)
is
Dx Dy Dz
x= , y= , z= ,
D D D
where
a1 b1 c1 d1 b1 c1 a1 d1 c1
D = a2 b2 c2 , D x = d 2 b2 c2 , Dy = a2 d2 c2 ,
a3 b3 c3 d3 b3 c3 a3 d3 c3
a1 b1 d1
D z = a2 b2 d2 .
a3 b3 d3
Proof. Left as an exercise to the reader.
Note 2.3.1 Since determinants are used to solve a system a linear equations by means of Cramer’s
rule, this method is also referred as determinant method.
Exercise 2.3.1 Using determinant method, solve the following system of linear equations
2x + 3y + 3z = 1
3x − 4y + z = 3
x − z = 0.
23
Theorem 2.4.2 If A is a triangular matrix, then
Here
det(I3 ) = 1 × 1 × 1 = 1.
det(A) = 0.
5 10 0
Example 2.4.5 Consider the matrix A = 0 0 0 . Here all entries along the 2nd row
2 6 5
of A are 0 (zero) and its determinant is
det(A) = 0.
Theorem 2.4.4 If any row (or a column) in a square matrix A is a multiple of another row (or
column) of it, then det(A) = 0.
1 3 0
Example 2.4.6 Consider the matrix A = 3 9 −1 . Here elements of the 2nd column
7 21 1
st
are 3 times of the corresponding elements of the 1 column in matrix A and
det(A) = 1(9 − 9) = 0.
24
If the ratios of the corresponding elements of any two rows or columns are equal in a
square matrix A, then det(A) = 0
is
not the equivalent
statement of the above theorem. For example, in case of the matrix A =
3 9 0
1 3 0, elements of the 1st row are 3 times of the corresponding elements of the 2nd
17 21 1
0
row. But here the ratio of the third elements of these two rows is , which is in indeterminant
0
form. In this case also
det(A) = 1(9 − 9) = 0.
Corollary 2.4.3 If a square matrix A has two identical rows or columns, then det(A) = 0.
1 3 10
Example 2.4.7 Consider the matrix A = 1 3 10 . Here 1st row and 2nd row in matrix
7 21 1
A are identical and
det(A) = 0.
a a x
Problem 2.4.1 Let ∆ = det m m m. The roots of the equation ∆ = 0 are
b x b
(a) x = a, m
(b) x = b, m
(c) x = a, b
(d) None.
Answer (b) x = a, b.
Theorem 2.4.5 If two rows or two columns of a square matrix are interchanged, then the deter-
minant of the new matrix will be the negative of the old one.
Example 2.4.8
a11 a12 a13 a11 a12 a13
a21 a22 a23 = − a31 a32 a33
a31 a32 a33 a21 a22 a23
or
a11 a12 a13 a12 a11 a13
a21 a22 a23 = − a22 a21 a23
a31 a32 a33 a32 a31 a33
Theorem 2.4.6 If each element of a row or of a column in a square matrix A are multiplied by
a constant k, then the determinant of the new matrix A′ will be equal to k · det(A).
Example 2.4.9
ka11 ka12 ka13 a11 a12 a13
a21 a22 a23 = k a21 a22 a23
a31 a32 a33 a31 a32 a33
25
or
ka11 a12 a13 a11 a12 a13
ka21 a22 a23 = k a21 a22 a23
ka31 a32 a33 a31 a32 a33
Theorem 2.4.7 If each element of a row (column) in a square matrix A are increased or de-
creased by the equal multiple of the corresponding elements of another row (or column), then the
determinant of the new matrix A′ remains same as det(A).
Example 2.4.10
a11 + ka12 a12 a13 a11 a12 a13
a21 + ka22 a22 a23 = a21 a22 a23
a31 + ka32 a32 a33 a31 a32 a33
or
a11 a12 − ka11 a13 a11 a12 a13
a21 a22 − ka21 a23 = a21 a22 a23
a31 a32 − ka31 a33 a31 a32 a33
α α x
Problem 2.4.2 For det β
β β = 0, find x.
θ x θ
Solution Here
α α x
det β
β β = 0
θ x θ
implies that
0 α x
det 0 β β = 0, c′1 = c1 − c2
θ−x x θ
0 α x
⇒(θ − x) · det 0
β β = 0
1 x θ
0 α x
⇒β(θ − x) · det 0
1 1 = 0
1 x θ
⇒(θ − x) · (−1)3+1 · 1 · (α − x) = 0
⇒(θ − x)(α − x) = 0
∴ x = α, θ.
Answer x = α, θ.
Theorem 2.4.8 If each entry of a row (column) in a square matrix A of order n × n can be
expressed as the sum of the entries of corresponding rows (columns) of two square matrices A′ and
A′′ with size n × n, then det(A) can be expressed as the sum of det(A′ ) and det(A′′ ).
Example 2.4.11
26
or
a11 + k1 a12 + k2a13 + k3 a11 a12 a13 k1 k2 k3
a21 a22 a23 = a21 a22 a23 + a21 a22 a23
a31 a32 a33 a31 a32 a33 a31 a32 a33
f (x) ϕ(x)
Problem 2.4.3 Let F (x) = det . Then establish the equality
g(x) ϕ(x)
F (x + h) − F (x) = ∆1 + ∆2 ,
f (x + h) − f (x) ϕ(x + h) f (x) ϕ(x + h) − ϕ(x)
whenever ∆1 = det and ∆2 = det .
g(x + h) − g(x) ϕ(x + h) g(x) ϕ(x + h) − ϕ(x)
1 4 x 22 4 x
Problem 2.4.4 Solve det 2 5 8 = det 26 5 8 .
3 6 9 30 6 9
Solution Here
1 4 x 22 4 x
det 2
5 8 = det 26
5 8
3 6 9 30 6 9
1 4 x 22 4 x
⇒ det 2
5 8 − det 26
5 8 = 0
3 6 9 30 6 9
1 − 22 4 x −21 4 x
⇒ det 2 − 26
5 8 = 0 ⇒ det −24
5 8 = 0
3 − 30 6 9 −27 6 9
7 4 x 7 4 x
⇒ (−3) · det 8
5 8 = 0 ⇒ det 8
5 8 = 0
9 6 9 9 6 9
Sice the determinant is zero, 1st column and 3rd column are identical. Therefore x = 7.
Answer 7.
Theorem 2.4.9 If A and B are two square matrices with same order, then
det(AB) = det(A)det(B).
Above Theorem 2.4.9 reflects the fact that Determinant of the product of two square matrices
with same order is equal to the product of their individual determinant. That means, determinant
is a multiplicative function.
Definition 2.4.1 The determinant of a Vandermonde matrix of order n is called the Vander-
monde determinant of order n.
Answer x = 0, 2.
27
2.5 Construction of a Square Matrix with Given Order and
Determinant
Here we study how can a square matrix of order n × n having determinant ∆ be formed.
Case 01 : The minor of the remaining element is non zero. Just write a in place of
the remaining entry, expand the determinant, set it equal to ∆ and solve the equation for a.
There you have your matrix.
Case 02 : The minor of the remaining element is zero. Change one of the other values
that occur in the calculation of the minor so that the minor becomes non zero and then follow
Case 01.
Question 2.5.1 How many 1 × 1 square matrix are there having ∆ as its determinant?
Answer Unique.
Question 2.6.1 Can AB be equal to I without being BA equal to I? Explain with example.
1 1
1 3 −1
Answer Yes. Consider that A = and B = −1
0. Here AB = I, but
1 1 0
−3 1
BA ̸= I.
Note 2.6.1 Some similar pairs of matrices satisfying the answer of Question 2.6.1
are as follows.
−2
0 1
2 2
1
0 0 1 1 −1 −1 1 2 0 3 1
, 0 0 ;
, 1 1 ; , − ;
1 1 0 −1 2 1 3 4 0 2 2
1 0 0 1 0 0
1 −2 2 −3
6 2 1 2 3 0
, −2 5 ; , −1 2 etc.
5 1 3 1 2 0
−1 2 0 0
28
Question 2.6.2 Why in the definition of inverse matrices both left product AB and right product
BA must be I?
Answer Because for some pairs of matrices A and B, it may happen that AB = I whereas
BA ̸= I. For instant, consider the pair of matrices
−3 −3
2 −3 4
A= and B = 3 2 .
−2 2 −3
4 3
Here AB = I, though BA ̸= I.
Proof. Let A be a square matrix. If possible, then let us suppose that B, C are two inverse matrices
of A. So
AB = BA = I and AC = CA = I.
Now,
B = BI
= B(AC)
= (BA)C; [as matrix multiplication is associative]
= IC = C.
Note 2.6.4 Sine the inverse of a matrix is unique, the inverse of a matrix A, if exists, is also
unique and it is denoted by A−1 .
(AB)−1 = B −1 A−1 .
29
Proof. Let A and B be two invertible matrices. So A−1 and B −1 exist.
Now,
(AB)(B −1 A−1 ) = A(BB −1 )A−1 = AIA−1 = (AI)A−1 = AA−1 = I.
Again,
(B −1 A−1 )(AB) = B −1 (A−1 A)B = B −1 IB = B −1 (IB) = B −1 B = I.
Hence (AB)−1 = B −1 A−1 .
Proof.
The next theorem known as determinant test for invertibility provides an important crite-
rion to determine whether a matrix is invertible.
30
2.6.2 Adjoint matrix
Let A = aij be square matrix and Cij be the corresponding cofactor of the entry aij . Then the
adjoint matrix of A is denoted by adj(A) and defined as
T
adj(A) = Cij = Cji .
That is, adjoint matrix of a given square matrix is the transpose of its cofactor matrix.
−2 0 1
Example 2.6.4 Consider the matrix A = 2 3 0 given in Example 2.6.3. The
−1 −3 4
adjoint matrix of A is
12 −3 −3
adj(A) = −8 −7 2 .
−3 −6 −6
Proof.
The next Theorem provides a technique to find the inverse of any invertible square matrix.
Proof.
4 3 10 17
Problem 2.6.1 If A = and AB = , then find the matrix B.
2 1 4 7
31
Proof. We have
det(AA−1 ) = det(In )
⇒ det(A)det(A−1 ) = 1 [as A and A−1 are of same of order]
1
∴ det(A−1 ) = .
det(A)
Solution Now,
det(λA) = λn det(A) = λn σ.
So
1
det((λA)−1 ) =
det(λA)
1
= n .
λ σ
Problem 2.6.3 Problem 34 of [2, Page 50 of Exercise 1.4]
det(AAT ) = det(In )
⇒det(A)det(AT ) = 1
⇒det(A)det(A) = 1; [by Theorem 2.4.1]
1
⇒det(A2 ) = [Why? Not Understanding]
det(A)
∴ det(A) = ±1 .
det(AAT ) = det(In )
⇒det(A)det(AT ) = 1
⇒det(A)det(A) = 1; [by Theorem 2.4.1]
⇒(det(A))2 = 1
∴ det(A) = ±1 .
det : T → R.
32
Solution Range(det) = {−1, 1}.
Problem 2.6.6 (Application of the Problem 2.6.5) Define an orthogonal matrix of order 2 × 2
that includes trigonometric identities cos θ and sin θ as its entry.
a b
Solution Let our required orthogonal matrix be A = . According to the definition of
c d
T
AA = I2 and so either det(A) = 1 or det(A) = −1.
orthogonal matrix,
a c
Here AT = and so
b d
2
T a + b2 ac + bd 1 0
AA = = I2 =
ac + bd c2 + d2 0 1
⇒ a2 + b2 = c2 + d2 = 1, ac + bd = 0 .
Now,
for any real number ρ, if we take c = ρ sin θ and d = −ρ cos θ or c = −ρ sin θ and d = ρ cos θ,
then ac + bd becomes 0 (zero). But in this problem c2 + d2 = 1. So we have to consider ρ = 1.
Thus either c = sin θ and d = − cos θ or c = sin θ and d = − cos θ.
33
Chapter 3
For better understanding and detail discussions of this chapter the reader is referred to Anton and
Rorres [2, Chapter 1], Lipschutz and Lipson [6, Chapter 3], Wildgerger [8].
a1 x 1 + a2 x 2 + a3 x 3 + · · · + an x n = b
is called a linear equation in the variables xi , where ai , b are constants for 1 ≤ i ≤ n and all ai
are not zero at a time.
Definition 3.1.2 (Homogeneous linear equation) [2, Page 2] A linear equation in n variables x1 ,
x2 , x3 , · · · , xn having the form
a1 x1 + a2 x2 + a3 x3 + · · · + an xn = 0
Definition 3.1.3 (Degenerate linear equation) [6, Page 62] A linear equation expressible in the
form
0x1 + 0x2 + 0x3 + · · · + 0xn = b
is called a degenerate linear equation in the variables xi .
The solution of a degenerate linear equation only depends on the value of b. If b ̸= 0, then
0x1 + 0x2 + 0x3 + · · · + 0xn = b has no solution. On the other hand, if b = 0, it has infinitely
many solutions. In fact, any n-tuples of Rn -space is a solution of the degenerate linear equation in
n-variables.
Exercise 3.1.1 Write down which of the following equations are linear and which are not.
√
(a) 2x2 + y − 3 = 0 (b) x − 5y 3 − 3 = 0 (c) x − 5y − 3 = 0
3
(d) x − 5y = 8 (e) x − 5y − 10 = 0 (f ) 3 − 5y − 103 = 0
x
x
(g) xy − 5y = 8 (h) − 5y − 10 = 0 (i) 3xy − 5y − 103 = 0.
y
34
3.1.2 System of linear equations
A system of linear equations is nothing but a finite set of linear equations.
Definition 3.1.4 (Linear system) [2, Page 2] A system of m linear equations in n unknown
variables x1 , x2 , x3 , · · · , xn is of the form
a11 x1 + a12 x2 + a13 x3 + · · · + a1n xn = b1
a21 x1 + a22 x2 + a23 x3 + · · · + a2n xn = b2
a31 x1 + a32 x2 + a33 x3 + · · · + a3n xn = b3
.. .. .. ... .. ..
. . . . .
am1 x1 + am2 x2 + am3 x3 + · · · + amn xn = bm
The the variables x1 , x2 , x3 , · · · , xn are called unknowns.
Exercise 3.1.2 Write down which of the following system of equations are System of linear
equations and which are not :
(a) 2x√+ y − 8 = 0 (b) 3x√+ 5y − 8 = 0 (c) x + 5y − 87 = 0
x+5=0 x + 5y = 8 x + 5y = 12
(d) xy + 5y − 9 = 0 (e) ex + 5y − 9 = 0 (f ) 2x + 5y − 9 = 0
5
x + 5y = 12 x + 5y = 12 + 5y = 12.
x
Definition 3.1.5 A solution of a linear system in n unknowns x1 , x2 , x3 , · · · , xn is the ordered
n-tuple
(x1 , x2 , x3 , · · · , xn ) = (s1 , s2 , s3 , · · · , sn )
for which the substitution x1 = s1 , x2 = s2 , x3 = s3 , · · · , xn = sn makes each equation a true
statement.
Definition 3.1.6 (Equivalent linear systems) [6, Page 63] Two linear systems are called equiv-
alent, if they have same solution set.
Every consistent linear system has zero, one or infinitely many solutions.
35
3.2.2 Matrix representation of a general linear system
The general linear system
can be represented as
a11 a12 a13 ··· a1n x1 b1
a21
a22 a23 ··· a2n x2 b2
a31
a32 a33 ··· a3n
x3 = b 3 .
.. .. .. ... .. .. ..
. . . . . .
am1 am2 am3 ··· amn xm bm
Definition 3.2.1 Let A⃗x = ⃗b be the matrix equation of a system of linear equations
Then the augmented matrix of the system is obtained by adjoining ⃗b to A as the last column as
a11 a12 a13 · · · a1n b1
a21 a22 a23 · · · a2n b2
A|⃗b = a31 a32 a33 · · · a3n b3 .
.. .. .. ... .. ..
. . . . .
am1 am2 am3 · · · amn bm
⃗ ⃗ = ⃗b.
The matrix A|b is called the augmented matrix of the system of the linear equation AX
h i
Note 3.2.1 Vertical partition line | in A|⃗b is optional [2, Page 34].
36
3.3 Elementary Row Operations
3.3.1 Objective of algebraic operation on a linear system
[2, Page 7] Each row in augmented matrix corresponds to the equations in the associated linear
system. To solve a linear system, algebraic operations are performed in such a way so that solution
set of the system remains unaltered and the system becomes simpler.
(ii) Multiply a row through by a nonzero constant: kri → ri for a nonzero constant k.
(iii) Add a constant times one row to another: kri + rj → rj for a nonzero constant k.
Answer Each row in augmented matrix corresponds to the equations in the associated linear
system, but a column does not do so. Moreover, if column operation is defined, then coefficient of
x will be added to the coefficient of y, which can not be. Because then linear system will be broken
down.
can be represented as
a1 b1 c1 x d1
a2 b2 c2 y = d2
a3 b3 c3 z d3
37
and solution of system (3.2) is
−1
x a1 b1 c1 d1
y = a2 b2 c2 d2
z a3 b3 c3 d3
Problem 3.4.1 Using the concept of inverse matrix, solve the following system of linear equations
2x + 3y = 1
3x − 4y = 3
Solution Given system of linear equations is
)
2x + 3y = 1
(3.3)
3x − 4y = 3
Exercise 3.4.1 Using the concept of inverse matrix, solve the following system of linear equations
2x + 3y + 3z = 1
3x − 4y + z = 3
x − z = 0.
38
Pivot variable and free variables
See [6, Page 68].
Note 3.4.1 Forward elimination part of Gaussian elimination tells that whether the linear system
is consistent or inconsistent. If the system is inconsistent, then the backward elimination part should
not be applied.
by Gaussian elimination.
x − 3y − 2z = 6
2x − 4y − 3z = 4
−3x + 6y − 8z = −5.
by Gaussian elimination.
Solution x = 1, y = −3, x = 2.
x+y−z =8
2x − y + 3z = 9
2x + 5y + z = 8
x + 2y − 3z = 1
2x + 5y − 8z = 4
3x + 8y − 13z = 7.
by Gaussian elimination.
Problem 3.4.6 [2, Problem 25 & 26 of Exercise 1.2] Determine the value of a for which the
following system has no solutions, exactly one solution, or infinitely many solutions.
39
Solution (a) For a = −4 the system has no solution. For a ̸= −4 and a ̸= 4, it has unique
solution. For a = 4, it has infinitely many solutions.
√ √ √
(b) For a = 2 or a = − 2 the system has no solution. For a ̸= ± 2, it has unique solution.
There is no value of a for which the system has infinitely many solutions.
(c) There is no such value of a, for which the system has no solution. For a ̸= 3, it has unique
solution. For a = 3, it has infinitely many solutions.
Problem 3.4.7 [2, Problem 27 & 28 of Exercise 1.2] What condition, if any, must a, b, and c
satisfy for the following linear system to be consisitent.
(a) x + 3y − z = a (b) x + 3y + z = a
x + y + 2z = b −x − 2y + z = b
2y − 3z = c 3x + 7y − z = c.
1 3 −1
a
3 a b
0 1 − −
2 2 2
0 0 0 −a + b + c
3.5 Application
3.5.1 Application in network flow
Problem 3.5.1 Example 1 [2, Page 84].
40
Chapter 4
For detail explanations and solving more problems readers are referred to Anton and Rorres [2],
Lipschutz and Lipson [6], Abdur Rahman[7].
(b) Reading top to bottom, the leading entries are to the right of the previous one.
∼
v
Problem 4.1.2 Using the row reduction solve the system of equations
2x + y + 3z = 1
x−y+z =3
x − z + 2z = 0.
41
Solution Augmented matrix of the given system is
2 1 3 1 1
1 −1 1 3 3
1 −1 2 0 0
1 2 2 −1 0
r2′ = r1 − r2
∼ 0 0 −1 −2 0 ;
r3′ = 3r1 − r3
0 0 −2 −4 0
1 2 2 −1 0
r2′ = (−1)r1
∼ 0 0 1 2 0 ;
r3′ = (− 12 )r1
0 0 1 2 0
1 2 2 −1 0
∼ 0
0 1 2 0 ; r3′ = r2 − r3
0 0 0 0 0
1 2 0 −5 0
∼ 0
0 1 2 0 ; r1′ = r1 − 2r2
0 0 0 0 0
Note 4.1.1 [6, Remark 2 of Page 73] If a system has more than four unknowns and four equations,
then it may be more convenient to employ the matrix format to solve a linear system.
Definition 4.1.2 A matrix in row echelon form is said to be in reduced row echelon form, if
the following two conditions hold:
(ii) each column containing a leading 1 has zeros everywhere else in that column.
Note 4.1.2
42
Problem
4.2.2 [2, Example 4, Page 56] Employ the row operations to find the inverse of
1 2 3
2 5 3, if exists.
1 0 8
43
Chapter 5
Vector Spaces
For detail explanations, better understanding and solving more problems readers are referred to
Anton and Rorres [2], Lipschutz and Lipson [6].
5.1 Vectors in Rn
See [2, Chapter 2]
Theorem 5.1.1 Let A and B be two points on a line l. Then the vector equation of l is
−→
A + λ direction vector AB
Note 5.1.1 Vector equation of a line is not unique, it is highly non unique.
Example 5.1.1 Vector equation of a line passing through the points (2, 0) and (6, 2) is
Example 5.1.2 Vector equation of the line passing through the points A(3, 1, 1) and B(2, 2, −1)
is
−→
(3, 1, 1) + λ direction vectorBA , where λ is a parameter
= (3, 1, 1) + λ(3 − 2, 1 − 2, 1 − (−1))
= (3, 1, 1) + λ(−1, −1, 2).
44
Answer Infinitely many.
Question 5.1.2 How many parameters are involved in the vcector equation of a straight line?
Answer One.
Example 5.1.3 Vector equation of the plane containing the points (1, −1, 0), (2, 1, 4) and (1, −1, 9)
is
(x, y, z) = (1, −1, 0) + λ[(2, 1, 4) − (1, −1, 0)] + η[(1, −1, 9) − (1, −1, 0)],
where λ and η are two parameters.
Theorem 5.1.2 If →
−
v = (v1 , v2 , v3 , · · · , vn ) is a vector and α is a scalar in Rn ,
(i) ||→
−
v || ≥ 0.
→
− →
−
(ii) || V || = 0 ⇔ V = 0.
→
− →
−
(iii) ||α V || = |α|||α V ||.
45
5.1.6 Inequalities
Theorem 5.1.3 (Cauchy-Schwarz inequality) If →
−
u = (u1 , u2 , u3 , · · · , un ) and →
−
v = (v1 , v2 , v3 , · · · , vn )
n
be two vectors in R , then
→
−u ·→−
v ≤ ||→
−
u || ||→
−
v ||.
Theorem 5.1.5 (Triangle inequality for distances) If u, v, w ∈ Rn are any three points, then
d (u, v) ≤ d(u, w) + d(w, v).
Definition 5.2.2 (Addition) Let V be a nonempty set and ⃗u, ⃗v ∈ V . Then addition or sum of
⃗u and ⃗v , denoted by ⃗u + ⃗v , is a function from V × V to V that assigns every (⃗u, ⃗v ) ∈ V × V to a
unique →−u +→ −
v ∈V.
Definition 5.2.3 (Vector Spaces) [2, Page 184] Let V be a nonempty set and F be a field with
underlying set F . Suppose that scalar multiplication and addition are defined on V . Then V
⃗ ∈ V and
together with these two operations is called a vector space over the field F, if for ⃗u, ⃗v , w
α, β, 1 ∈ F , the undermentioned axioms are obeyed:
(i) ⃗u + ⃗v = ⃗v + ⃗u
(ii) ⃗u + (⃗v + w)
⃗ = (⃗u + ⃗v ) + w
⃗
(iii) there exists a ⃗0 ∈ V such that ⃗u + ⃗0 = ⃗0 + ⃗u = ⃗u
⃗ = (−u)
(iv) for each ⃗u ∈ V , there exists a −⃗u such that ⃗u + (−u) ⃗ + ⃗u = 0
Example 5.2.2 Set of all 2 × 2 matrices form a vector space with addition and scalar multipli-
cation.
Example 5.2.3 Let V be the set of all real-valued functions defined on R. Then V forms a vector
space with addition and scalar multiplication of real-valued functions.
46
5.2.1 Real vector space
Scalars k ∈ R come from real field [2, Page 184, side note]. That means, here the scalars are
numbers.
47
if for all ⃗v ∈ V ,
⃗v = α1 v⃗1 + α2 v⃗2 + α3 v⃗3 + · · · + αn v⃗n
for some numbers α1 , α2 , α3 , · · · , αn . If S is a spanning set of V , then it is said that S spans V .
1 0 0 2 3
Example 5.4.1 0 , 1 , 0 , 1 spans R3 . For example, 1
0 0 1 −5 4
1 0 2 3 0
Example 5.4.2 0 , 1 , 5 , −1 does not span R3 . For example, 0 can not
0 0 0 0 1
be written as a linear combination of the vectors of the given set.
1 2 −1
Problem 5.4.1 Does 2 , 0 , 4 span R3 ? Why?
−1 1 3
y1
Solution Let y2 be any vector of R3 . Yes. Unique solution.
y3
Definition 5.4.2 (Basis of a vector space) [2, Page 214] Let S = {v⃗1 , v⃗2 , v⃗3 , . . . , v⃗n } be a set of
vectors in a finite dimensional vector space V . Then S is called a basis for V if the following two
conditions hold :
(i) S spans V
Theorem 5.4.1 [2, Theorem 4.5.1 of Page 221] All bases of a finite-dimensional vector space
have the equal number of vectors.
5.4.2 Dimension
Definition 5.4.3 [2, Page 222] Let the set of vectors S be a basis of a finite dimensional vector
space V . Then the dimension of V , denoted by dim(V ), is the number of vectors of S. Zero
vector space is defined to have the dimension zero.
Problem 5.4.2 Define basis of a vector space. Find a basis for the solution space of the linear
system of equations
x + 2y + 2z − s + 3t = 0
x + 2y + 3z + s + t = 0
3x + 6y + 8z + s + 5t = 0
48
system is
1 2 2 −1
3 0
1 2 3 11 0
3 6 8 15 0
1 2 2 −1 3 0
r2′ = r1 − r2
∼ 0 0 −1 −2 2 0 ;
r3′ = 3r1 − r3
0 0 −2 −4 4 0
1 2 2 −1 3 0
r2′ = (−1)r1
∼ 0 0 1 2 −2 0 ;
r3′ = (− 12 )r1
0 0 1 2 −2 0
1 2 2 −1 3 0
∼ 0 0 1 2 −2 0 ; r3′ = r2 − r3
0 0 0 0 0 0
1 2 0 −5 7 0
∼ 0 0 1 2 −2 0 ; r1′ = r1 − 2r2
0 0 0 0 0 0
which is the reduced row echelon form of the corresponding augmented matrix of the given system.
Here y, s and t are the 3 independent variables. Say y = λ, s = µ and t = η. Then
So
Let →
−
v1 = (−2, 1, 0, 0, 0), →
−
v2 = (5, 0, −2, 1, 0) and →
−v3 = (−7, 0, 2, 0, 1). Thus the solution space is
spanned by the vectors → −v1 , →
−
v2 and →
−
v3 . By inspection, these vectors are linearly independent since
→
−
λ→
−
v1 + µ→
−
v2 + η →
−
v3 = 0 ⇒ λ = µ = η = 0.
Therefore →
−
v1 , →
−
v2 and →
−
v3 form a basis for the solution space and the dimension of the solution space
is 3.
Definition 5.5.2 (Column space) [2, Page 237] Let A be a m × n matrix. Then the subspace
of Rm spanned by the column vectors of A is called the column space of A and it is denoted by
col(A).
Definition 5.5.3 (Null space) [2, Page 237] Let A be a m × n matrix. Then the solution space
of the homogeneous linear system
AX⃗ = ⃗0
is called the null space of A and it is denoted by null(A). Null space is a subspace of Rn .
49
5.6 Rank and Nullity of a Matrix
For details see [2, Section 4.8].
Definition 5.6.1 [2, Section 4.8] The common dimension of the row space and column space of
a matrix A is called the rank of A. It is denoted by rank(A).
Definition 5.6.2 [2, Section 4.8] The dimension of the null space of a matrix A is called the
nullity of A. It is denoted by nullity(A).
Theorem 5.6.1 (Dimension theorem for matrices) [2, Theorem 4.8.2] If A is a matrix with n
columns, then rank(A) + nullity(A) = n.
Solution
50
Chapter 6
For detail explanations and solving more problems readers are referred to Anton and Rorres [2],
Lipschutz and Lipson [6], Abdur Rahman[7].
51
Chapter 7
For detail explanations, better understanding and solving more problems readers are referred to
Anton and Rorres [2], Lipschutz and Lipson [6], Wildgerger [8].
7.1 Eigenvalues
7.1.1 Eigenvalues of a matrix
Definition 7.1.1 (Eigenvalues of a matrix) [2, Page 291] Let A be a square matrix of order n × n
and ⃗v ∈ Rn such that ⃗v ̸= 0 and
det(λI − A) = 0.
52
Note 7.1.2 A matrix A of order n × n defined over the real field may have complex eigenvalues.
(i) λ is an eigenvalue of A.
(iv) there exists a nonzero vector ⃗v belonging to Rn such that A⃗v = λ⃗v .
Definition 7.1.5 (Eigenspace and its basis) [2, Page 295] Let A be a n × n matrix, λ be an
eigenvalue and ⃗v be the corresponding eigenvector of A. Then the solution space of the linear
system
(λI − A)⃗v = ⃗0
is called the eigenspace of A corresponding to λ and the basis of this solution space is called the
corresponding basis for the eigenspace.
4 −2
Problem 7.1.2 Find the eigenvalues and eigenvectors of the matrix .
2 6
−1 3
Problem 7.1.3 [2, Example 6] Find bases for the eigenspaces of the matrix . Give the
2 0
geometrical interpretation of the eigenvalues and eigenvectors of this matrix with the help of a
sketch.
0 0 −2
Problem 7.1.4 [2, Example 7] Find the eigenvalues and eigenvectors of the matrix 1 2 1 .
1 0 3
0 0 −2
Or, find the bases for the eigenspaces of the matrix 1 2 1 .
1 0 3
cos θ − sin θ
Problem 7.1.5 [2, Problem 1(a), Page 343] Show that if 0 < θ < π, then has
sin θ cos θ
no real eigenvalues and consequently no real eigenvectors.
53
Problem 7.1.6 [2, Like Problem 27 of Exerecise Set 5.1, Page 301] Find the matrix A such that
its eigenvalues are 2, 4, 6 and corresponding eigenvectors are
1 1 0
1 , 0 , 1
0 1 1
Solution Substitution of the given eigenvectors ⃗v and the corresponding eigenvalues λ into A⃗v =
λ⃗v yields the following equations
1 1 2
A 1 = 2 1 = 2 ,
0 0 0
1 1 4
A 0 = 4 0 = 0
1 1 4
and
0 0 0
A 1 = 6 1 = 6 .
1 1 6
Above three equations can be written together as
1 1 1 2 4 0
A 1
0 1 = 2
0 6
0 1 0 0 4 6
−1
2 4 0 1 1 1
⇒ A = 2 0 6 1 0 1 .
0 4 6 0 1 0
The next theorem describes a relationship between eigenvalues and the invertibility of a matrix.
7.2 Diagonalization
For details go through Anton and Rorres [2, Section 5.2].
54
Now,
= AI = A.
Note 7.2.1 A and P −1 AP are similar and have same determinant, rank, nullity, trace, charac-
teristic polynomial, eigenvalues etc [2, Table 1, Page 303].
55
0 0 −2
Problem 7.3.1 Verify the Cayley-Hamilton theorem for the matrix 1
2 1 .
1 0 3
1 0 0 0 0
0 2 0 0 0
0
Problem 7.3.2 Verify the Cayley-Hamilton theorem for the matrix 0 3 0 0 .
0 0 0 8 0
0 0 0 0 −1
7.3.2 Importance
The Cayley-Hamilton theorem helps to calculate the power of a matrix.
5
2 1 1
Problem 7.3.3 Find 1 1 0 using the Cayley-Hamilton theorem.
1 0 1
5
2 1 1
Solution Let 1 1 0 . Eigenvalues of A are 0, 1, 3. So the corresponding characteristic
1 0 1
3 2
equation is λ − 4λ + 3λ = 0. Thus by the Cayley-Hamilton theorem,
Thus A4 = 4A3 −3A2 = 4(4A2 −3A)−3A2 = 13A2 −12A. Therefore A5 = 13A3 −12A2 = 40A2 −39A
5
2 0 0
Problem 7.3.4 Find 1 1 0 using the Cayley-Hamilton theorem.
1 0 −1
56
Chapter 8
Linear Transformation
For detail explanations and solving more problems readers are referred to Anton and Rorres [2,
Chapter 8], Lipschutz and Lipson [6].
8.1.3 Combination
0 1 1 0 0 −1
=
1 0 0 −1 1 0
Problem 8.1.1 [2, Like Problem 1, 2, 20, 21 of Exerecise Set 4.11, Page
287-288]
Find and sketch
3 1
the image of y = 2x + 1 under the multiplication of the matrix A = .
6 2
Solution Let the coordinates (x, y) on the line y = 2x + 1 be transformed to the point with
coordinates (x′ , y ′ ) under the multiplication by A. Then
′
x 3 1 x
′ =
y 6 2 y
′
x 3x + y 3x + y
⇒ = = .
y′ 6x + 2y 2(3x + y)
The points on the plane R2 maps to the line y = 2x under the described map.
→
− →−
Theorem 8.1.1 [2, Theorem 8.1.1] If T : V → W is a linear transformation, then T 0 = 0
and T (→
−
u −→
−
v ) = T (→
−
u ) − T (→
−
v ) for all →
−
u ,→
−
v ∈V.
57
Problem 8.1.2 [2, Example 10, Page 451] Consider the basis {(1, 1, 1), (1, 1, 0), (1, 0, 0)} for R3 .
Suppose that T : R3 → R2 is a linear transformation for which
T (1, 1, 1) = (1, 0), T (1, 1, 0) = (2, −1), T (1, 0, 0) = (4, 3).
Determine a formula for T (x, y, z), and then employ this formula to compute T (2, −3, 5).
Theorem 8.2.1 [2, Theorem 8.1.3] If T : V → W is a linear transformation, then ker(T ) and
R(T ) are subspaces of V and W respectively.
Theorem 8.3.1 (Dimension theorem for linear transformation) [2, Theorem 8.1.4] If T is a linear
transformation from a finite dimensional vector space V to a vector space W , then the range of T
is finite dimensional and rank(T ) + nullity(T ) = dim(V ).
Problem 8.3.1 Determine the kernel of the linear transformation defined by the matrix
1 0 4 −2
1 −1 3 0 .
1 1 5 −4
Hence find the dimension of the kernel space (or nullity of the transformation).
58
3 5
Find the matrix transformation T with respect to the bases B = , for R2 and B ′ =
1 2
1 −1 0
0 , 2 , 1 for R3 .
−1 2 2
59
Chapter 9
Quadratic Forms
For detailed explanations and solving more problems, readers are referred to Anton and Rorres [2,
Section 7.3 and 7.4].
9.2 Application
See Anton and Rorres [2, Section 7.4].
60
Part II
61
Chapter 10
For detail explanations and solving more problems readers are referred to Anton and Rorres [2],
Lipschutz and Lipson [6], Wildberger [8].
Note 10.1.1 Parametric equation of a line is not unique, it is highly non unique.
Example 10.1.1 Parametric equation of a line passing through the points (2, 0) and (6, 2) is
Example 10.1.2 Parametric equation of the line passing through the points A(3, 1, 1) and
B(2, 2, −1) is
−→
(3, 1, 1) + λ direction vectorBA , where λ is a parameter
= (3, 1, 1) + λ(3 − 2, 1 − 2, 1 − (−1))
= (3, 1, 1) + λ(−1, −1, 2).
62
Problem 10.2.1 Find the meet of the lines x − 2y = 2 and 5x + 4y = 20.
Solution
1 −2 x 2
=
5 4 y 20
−1
x 1 −2 2 1 4 2 2
⇒ = = .
y 5 4 20 14 −5 1 20
Problem 10.2.2 Find the meet of the lines
Example 10.3.1 Three points on the plane p : 2x + 6y + 3z = 6 are (0, 0, 2), (0, 1, 0) and (3, 0, 0).
Moreover,
when x = 0, then 2y + z = 2.
when y = 0, then 2x + 3z = 6.
when x = 0, then x + 3y = 3.
63
10.4 Lines in 3-dimensions
10.4.1 Lines in 3-dimensions
Lines in 3-dimensions are more complicated. Any line can be determined by two points on it.
Moreover, a line can be interpreted as the intersection of two planes.
Solution Let
(i) we can subtract the coordinates of any point (x0 , y0 , z0 ) lying on the line respectively from x,
y and z in the numerator and
(ii) we can place the corresponding components of any scalar multiple of the direction vector
in the corresponding denominators, as any scalar multiple of a direction vector is again a
direction vector.
64
10.4.3 Meet of a line a plane
If a line and a plane meet, then we get a point.
Problem 10.4.3 Find the meet of the line l : (−1, 2, −3)+λ(0, 4, 6) and the plane p : x−2y+3z =
6.
ax + by + cz = d
ex + f y + gz = h.
algebraically, then introduce a third plane z = λ and solve the new system. If
a b c
e f g = 0,
0 0 1.
x + 2y − z = −3
3x + 7y + 2z = 1.
Solution Let z = λ be another plane. Then the corresponding matrix equation of the system
x + 2y − z = −3
3x + 7y + 2z = 1
z=λ
is
1 2 −1 x −3
3 7 2 y = 1 .
0 0 1 z λ
65
1 2 −1 1 2 −1
Since 3 7 2 = 1(7 − 6) = 1 ̸= 0, the matrix 3
7 2 is invertible. Thus
0 0 1 0 0 1
−1
x 1 2 −1 −3
y = 3 7 2 1
z 0 0 1 λ
7 −2 11 −3 −23 + 11λ
= −3 1 −5 1 = 10 − 5λ
0 0 1 λ λ
Therefore meet of the given two planes is the line with parametric equations
That is, if we add any multiple of this vector with a solution of the given two equations of planes,
the solution remain same.
x + 3y − 2z = 2
2x + 6y − 5z = 3.
Solution Let z = λ be another plane. Then the corresponding matrix equation of the system
x + 3y − 2z = 2
2x + 6y − 5z = 3
z=λ
is
1 3−2 x 2
2 6−5 y = 3 .
0 0 1 z λ
1 3 −2 1 3 −2
Since 2 6 −5 = 0, the matrix 2 6 −5 is not invertible. Thus we introduce
0 0 1 0 0 1
another plane y = λ and now the new system becomes
x + 3y − 2z = 2
2x + 6y − 5z = 3
y=λ
66
is
1 3 −2 x 2
2 6 −5 y = 3 .
0 1 0 z λ
1 3 −2 1 3 −2
Since 2 6 −5 = 1 ̸= 0, the matrix 2 6 −5 is invertible. Thus
0 1 0 0 1 0
−1
x 1 3 −2 2
y = 2 6 −5 3
z 0 1 0 λ
5 −2 −3 2 4 − 3λ
= 0 0 1 3 = λ
2 −1 0 λ 1
Therefore meet of the given two planes is the line with parametric equations
(x, y, z) = (4 − 3λ, λ, 1) = (4, 0, 1) + λ(−3, 1, 0),
x−4
which has the cartesian equation = y, z = 1.
−3
Example 10.6.1 Parametric equation of the plane containing the points (1, −1, 0), (2, 1, 4) and
(1, −1, 9) is
(x, y, z) = (1, −1, 0) + λ[(2, 1, 4) − (1, −1, 0)] + η[(1, −1, 9) − (1, −1, 0)],
where λ and η are two parameters.
67
Problem 10.6.1 Obtain the cartesian equation of the plane
x−1 1 0
y+1 2 0 =0
z 4 9
⇒ (−1)3+3 9[2(x − 1) − (y + 1)] = 0 ∴ 2x − y = 3.
Solution Clearly, A = (0, 0, 2), B = (0, 6, 0) and C = (3, 0, 0) lie on the given plane.
68
Bibliography
[2] H. Anton and C. Rorres, Elementary Linear Algebra (Applications Version), Eleventh ed., Jhon
Wiley & Sons, 2020-21.
[3] G. Hadley, Linear Algebra, Addison-Wesley World Student Series ed., Addison Wesley Publish-
ing Company, INC, 1977.
[4] A. G. Hamilton, Linear Algebra : An introduction with concurernt example, First ed., Cambridge
University Press, Cambridge, 1989.
[5] R. Larson and D. C. Falvo, Elementary Linear Algebra, Sixth ed., Houghton Mifflin Harcourt
Publishing Company, Boston, Newyork, 2009.
[6] S. Lipschutz and M. L. Lipson, Theory and problems of linear Algebra, Third ed., Schaum’s
Outline Series, McGraw-Hill, 2013-2014.
[7] M. Abdur Rahman, College Linear Algebra: Theory of Matrices with Applications, Sixth
(Reprint) ed., Nahar Book Depot & Publications, Dhaka, 2012.
69