On Determinat of Non Square Matrices
On Determinat of Non Square Matrices
By
Supervisors
At Hebron, Palestine.
2021
On Determinant Of Non-Square Matrices
By
i
اإلهداء
إلى من أشتاق إليه بكل جوارحي ...إلى أبي الذي فارقنا بجسده ،ولكن روحه ما زالت تُرفرف في
سماء حياتي.
إلى نبع المحبة واإليثار والكرم ...إلى من أمدَّتني بالنصح واإلرشاد والدعاء ...أمي أطال هللا في
ii
Acknowledgments
First of all, I thank my God for all the blessing he bestowed on me and continues
to bestow on me.
I would also like to express my gratitude and appreciation for my sister Sawsan
Dweak for her help, patience, encouragement, supporting this project.
My thanks are extended to Dr. Bassam Manasrah and Dr. Iyad Alhribat for
serving in my committee.
I thank my family members for the support and encouragement and for their
effective cooperation and care in preparing this thesis.
iii
Declaration
The work provided in this thesis, unless otherwise referenced, is the result of the
researcher's work, and has not been submitted elsewhere for any other degree or
qualification.
Signature:
Date :
iv
Contents
Abstract iiv
Preface 1
1 Preliminaries 3
1. Matrices and matrix operations 3
2. Properties of algebraic operations on matrices 4
3. Determinants of Matrices 5
4. Cofactor Expansion and Adjoint 7
5. The Inverse of a Matrix 8
6. Linear Systems 9
7. Gaussian elimination, Cramer's rule 11
8. Vectors in the Euclidean space 𝑅 𝑛 13
9. Rank and Nullity 14
2 Determinant 18
1. Determinant function 18
2. Minors and cofactors 21
3. Determinant Using Cofactor Expansion 23
4. Effect of elementary row operations on determinates 26
5. Radic's determinant 33
3 Properties for non-square determinant 42
1. Properties for determinant of a rectangular matrix. 42
2. How determinant is affected by operations on columns 51
4 Applications of Determinant to Area of Polygons in 𝑹 𝟐 58
1. Areas of certain polygons in connection with determinant of 58
rectangular matrices.
2. Some properties of the determinant and their geometric 63
interpretation
5 Inverse for non-square matrix 74
1. Right inverse or left inverse of a matrix 74
2. Properties for inverse and adjoint of non- square matrices 78
3. Pseudo inverse of non-square matrices 84
v
6 Applications to linear systems of equations 𝑨 𝑿 = 𝑩 87
1. Solving a linear system Using pseudo inverse 87
2. Using adjoint to solve a system of linear equations 91
3. Cramer's rule for nonsingular 𝑚 × 𝑛 matrices 94
4. Particular Cases and Examples 99
References 105
vi
Abstract
We study the definition of the determinant of a non-square matrix, using
cofactor definition and Radic definition, and we proved that they are identical by
proving the uniqueness of the determinant function that satisfies the four
characterizing properties of determinant function.
We also study the connection between the area of any polygon in the Cartesian
plane and determinant function for 2 × 𝑛 matrices. We used several methods to
find the mathematical isotope and prove the properties for inverse and adjoint for
a matrix as well as solving systems of equations in several ways.
vii
الملخص
قمنا بدراسة تعريف محدد المصفوفة غير المربعة باستخدام تعريف العامل المساعد وتعريف رادك ،
وأثبتنا أنهما متطابقان من خالل إثبات أن كالً منهما يحقق الشروط االربع للدالة المحددة.
ضا العالقة بين مساحة أي مضلع في المستوى الديكارتي والدالة المحددة لمصفوفات من
درسنا أي ً
𝑛×2 الرتبة
و أستخدمنا عدة طرق إليجاد النظير الرياضي وإثبات خصائص معكوس المصفوفة ومضاد
viii
Preface
In the books of a linear algebra we studied the concept of matrix and its types,
and we studied the definition of the determinant of the square matrix, its
properties and its applications. Here we will study the definition of the
determinant of the non-square matrix, its verified properties and its applications in
finding the area of polygons and finding solutions to the system of linear
equations.
In chapter (1) we introduce the basic results and definitions which shall be
needed in the following chapters. The topics include results about matrices and
matrix operations, properties of algebraic operations on matrices, Determinants of
Matrices, The Inverse of a Matrix, Cofactor Expansion, Adjoint of Matrix, Linear
Systems, Reduced Row-Echelon Form, Gauss-Jordan reduction, Cramer's rule,
Rank and Nullity and Vectors in the Euclidean space 𝑅𝑛 . This chapter is
absolutely fundamental. The results have been stated without proofs, for theory
may be looked in any text book in linear algebra. A reader who is familiar with
these topics may this chapter and refer to it only when necessary.
1
Chapter (3) we study Radic definition for determinant of a rectangular matrix
in more detailed way. We present new identities for the determinant of a
rectangular matrix. We develop some important properties of this determinant.
We generalize several classical important determinant identities, and description
how the determinant is affected by operation on columns, such as interchanging
columns, reversing columns or decomposing a single column.
Chapter (5) we will study existence of inverses for non-square matrices. Also,
we compute an inverse of a rectangular matrix using solution of a linear system
and an adjoint of matrices. In section (2) we study some important properties for
inverse and adjoint of non-square matrices. In section (3) we discuss Pseudo
inverse method which gives an inverse of matrices.
Chapter (6) we will discuss some results concerning the solutions of a linear
system 𝐴𝑥 = 𝑏 using inverses as well as the pseudo-inverse and adjoint of a
rectangular 𝑚 × 𝑛 matrix 𝐴, and General solution theorem. In section (3) we
study want to generalize this method of Cramer's for an 𝑚 < 𝑛 system of linear
equations. Finally in section (4) we study shall consider some particular cases and
examples to illustrate the results of what we have done in the previous sections
especially in applying pseudo inverse to some certain examples.
2
Chapter One
Preliminaries
This chapter contains some definitions and basic results about matrices and
matrix operations, properties of algebraic operations on matrices, Determinants of
Matrices, The Inverse of a Matrix, Cofactor Expansion, Adjoint of Matrix, Linear
Systems, Reduced Row-Echelon Form, Gauss-Jordan reduction, Cramer's rule,
Rank and Nullity and Vectors in the Euclidean space 𝑅𝑛 .
𝑎1𝑗
𝑎2𝑗
The 𝑗𝑡ℎ column of 𝐴 is 𝑎3𝑗 (1 ≤ 𝑗 ≤ 𝑛).
⋮
[𝑎𝑚𝑗 ]
3
Definition 1.1.3 [ 2 , 𝑝. 27 ]. Two matrices are defined to be equal if they have the
same size and their corresponding entries are equal.
Definition 1.1.5 [ 2 , 𝑝. 14 ].
a. (Matrix Addition) If 𝐴 = (𝑎𝑖𝑗 ) and 𝐵 = (𝑏𝑖𝑗 ) are 𝑚 × 𝑛 matrices, then the
sum of 𝐴 and 𝐵 is the 𝑚 × 𝑛 matrix 𝐶 = (𝑐𝑖𝑗 ), defined by 𝑐𝑖𝑗 = 𝑎𝑖𝑗 + 𝑏𝑖𝑗
( 1 ≤ 𝑖 ≤ 𝑚, 1 ≤ 𝑗 ≤ 𝑛).
Theorem 1.2.1 [ 9 , 𝑝. 35 ].
Let 𝐴 , 𝐵 , 𝐶 𝑎𝑛𝑑 𝐷 be an 𝑚 × 𝑛 matrices. and 𝑟 and 𝑠 are real numbers, then
(1) 𝐴 + 𝐵 = 𝐵 + 𝐴.
(2) 𝐴 + (𝐵 + 𝐶 ) = (𝐴 + 𝐵) + 𝐶 .
(3) There is a unique 𝑚 × 𝑛 matrix 𝑂 such that 𝐴 + 𝑂 = 𝐴 for any 𝑚 × 𝑛
matrix 𝐴. The matrix 𝑂 is called the 𝑚 × 𝑛 additive identity or zero matrix.
(4) For each 𝑚 × 𝑛 matrix 𝐴, there is a unique 𝑚 × 𝑛 matrix 𝐷 such that
𝐴 + 𝐷 = 𝑂. (1)
4
We shall write 𝐷 as −𝐴, so that (1) can be written as
𝐴 + (−𝐴) = 𝑂.
The matrix – 𝐴 is called the additive inverse or the negative of 𝐴.
(5) If 𝐴 , 𝐵 and 𝐶 are of the appropriate sizes, then 𝐴(𝐵 𝐶 ) = (𝐴 𝐵) 𝐶 .
(6) If 𝐴 , 𝐵 and 𝐶 are of the appropriate sizes, then 𝐴(𝐵 + 𝐶 ) = 𝐴 𝐵 + 𝐴 𝐶 .
(7) If 𝐴 , 𝐵 and 𝐶 are of the appropriate sizes, then (𝐴 + 𝐵 )𝐶 = 𝐴 𝐶 + 𝐵 𝐶 .
(8) 𝑟(𝑠 𝐴) = (𝑟 𝑠 )𝐴.
(9)( 𝑟 + 𝑠 )𝐴 = 𝑟 𝐴 + 𝑠 𝐴.
(10) 𝑟( 𝐴 + 𝐵 ) = 𝑟 𝐴 + 𝑟 𝐵.
(11) 𝐴 ( 𝑟 𝐵 ) = 𝑟 ( 𝐴 𝐵 ) = ( 𝑟𝐴 )𝐵.
For a proof for these properties [9].
(b) ( 𝐴 + 𝐵 )𝑇 = 𝐴𝑇 + 𝐵𝑇 and ( 𝐴 − 𝐵 )𝑇 = 𝐴𝑇 − 𝐵𝑇 .
(c) ( 𝐴 𝐵 )𝑇 = 𝐵𝑇 𝐴𝑇 .
(d) ( 𝑟 𝐴)𝑇 = 𝑟 𝐴𝑇 .
For a proof for these properties [9].
5
(4) If the matrix 𝐵 is the result of exchanging two rows of 𝐴, then the
determinant of 𝐵 is the negation of the determinant of 𝐴. (|𝐵| = −|𝐴|)
Theorem 1.3.2 [ 8]. A determinant function has the following four properties.
(a) The determinant of any matrix with an entire row of 0′ 𝑠 is 0.
(b) The determinant of any matrix with two identical rows is 0.
(c) If one row of a matrix is a scalar multiple of another row, then its
determinant is 0.
(d) If a scalar multiple of one row of a matrix is added to another row, then the
6
(g) If 𝐴 has a row (column) consisting of all zeros, then 𝑑𝑒𝑡(𝐴) = 0
For proofs of these properties [9].
Theorem 1.3.6 [9 , 𝑝. 93 ].
1. Let 𝐴 , 𝐵 and 𝐶 be an 𝑛 × 𝑛 matrices that differ only in a single row, say the
𝑟𝑡ℎ row, and assume that the 𝑟𝑡ℎ row of 𝐶 can be obtained by adding
corresponding entries in the 𝑟𝑡ℎ rows of 𝐴 and 𝐵. Then
𝑑𝑒𝑡( 𝐶) = 𝑑𝑒𝑡(𝐴) + 𝑑𝑒𝑡(𝐵)
The same result holds for columns.
𝑖𝑡ℎ row and 𝑗𝑡ℎ column of 𝐴. The determinant 𝑑𝑒𝑡(𝑀𝑖𝑗 ) is called the minor of 𝑎𝑖𝑗 .
The cofactor 𝐴𝑖𝑗 of 𝑎𝑖𝑗 is defined as 𝐴𝑖𝑗 = (−1)𝑖+𝑗 𝑑𝑒𝑡(𝑀𝑖𝑗 ).
7
𝐴11 𝐴21 ⋯ 𝐴𝑛1
𝐴 𝐴22 ⋯ 𝐴𝑛2
𝑎𝑑𝑗(𝐴) = [ 12 ].
⋮ ⋮ ⋱ ⋮
𝐴1𝑛 𝐴2𝑛 ⋯ 𝐴𝑛𝑛
𝐴 𝐵 = 𝐵𝐴 = 𝐼𝑛 .
The matrix 𝐵 is called an inverse of 𝐴. If there exists no such matrix 𝐵, then 𝐴 is called
singular (or noninvertible).
−1
(a) If 𝐴 is a nonsingular matrix, then 𝐴−1 is a nonsingular and (𝐴−1 ) =𝐴
8
Corollary 1.5.3 [ 9 , 𝑝. 100 ]. If 𝐴 is nonsingular, then 𝑑𝑒𝑡(𝐴) ≠ 0 and
1
det (𝐴−1 ) = det(𝐴)
A solution for the linear system is a sequence of 𝑛 real numbers s1, s2, ..., sn
which when substituted in the equations of the linear system all become true statements.
The principal question for this kind of systems is to find the set of solutions to this
system, that is to find all 𝑛-tuple (𝑥1 , 𝑥2 , . . . , 𝑥𝑛 ) that satisfy (1)
A general system of linear equations lies in one and only one of the following
categories
9
ii. The system has exactly one solution.
iii. The system has infinitely many solutions.
If there are fewer equations than variables in a linear system, then the system
either has no solution or it has infinitely many solutions.
The key idea in finding the solution of a linear system is to apply elementary
row operations
Theorem 1.6.2 [ 9 , 𝑝. 7 ]. If any finitely many operations of the following is (are) applied
to the linear system (1)
To get a new system. Then both systems have the same solution.
The operations listed in Theorem 1.6.2 are called elementary row operations.
10
The entries in the product 𝑨𝒙 are merely the left sides of the equations in (1).
Hence the linear system (1) can be written in matrix form as 𝑨𝒙 = 𝒃 .
The matrix 𝐴 is called the coefficient matrix of the linear system (1), and the
matrix
𝑎11 𝑎12 ⋯ 𝑎1𝑛 𝑏1
𝑎21 𝑎22 ⋯ 𝑎2𝑛 𝑏2
[ ⋮ ⋮ ⋱ ⋮ | ⋮ ],
𝑎𝑚1 𝑎𝑚2 ⋯ 𝑎𝑚𝑛 𝑏𝑚
(1). The augmented matrix of (1) will be written as [𝑨|𝒃]. Conversely, any matrix
with more than one column can be thought of as the augmented matrix of a linear
system. The coefficient and augmented matrices play key roles in solving linear
systems.
1- If a row does not consist entirely of zeros, then the first nonzero number in
the row is a 1. (We call this a leading 1.)
2- Any row that consists entirely of zeros is placed at the bottom of the matrix.
3- In any two successive rows that do not consist entirely of zeros, the leading 1
in the lower row occurs farther to the right than the leading 1 in the higher row.
4- Each column that contains a leading 1 has zeros everywhere else.
11
Any 𝑚 × 𝑛 matrix 𝐴 can be transformed into a reduced row echelon form 𝐴′
which is row equivalent to 𝐴, and this 𝐴′ is unique .
Step 3. The linear system that corresponds to the matrix in reduced row-echelon
form that has been obtained in step 2 has exactly the same solutions as the given
linear system. For each nonzero row of the matrix in reduced row-echelon form,
solve the corresponding equation for the unknown that corresponds to the leading
entry of the row. The rows consisting entirely of zeros can be ignored, since the
corresponding equation will be satisfied for any values of the unknowns.
equations in 𝑛 unknowns such that det(𝐴) ≠ 0, then the system has the unique
Where 𝐴𝑗 is the matrix obtained by replacing the entries in the 𝑗𝑡ℎ column of 𝐴
𝑏1
𝑏2
by the entries in the matrix 𝒃 = [ ] .
⋮
𝑏𝑛
12
1.8 Vectors in the Euclidean space 𝑹𝒏
In this section we state some basic definitions regarding the vector space 𝑅𝑛
𝑥1
A vector in the plane is a 2-vector 𝑢 = [𝑦 ], where 𝑥1 and 𝑦1 are real
1
Definition 1.8.2 [9 , 𝑝. 148 ]. (Norm of a vector) The length (also called magnitude or
called orthogonal if 𝑢. 𝑣 = 0.
13
component of 𝑢 along 𝑣. It is denoted by 𝑝𝑟𝑜𝑗𝑣 𝑢.
𝑢.𝑣
𝑢 − 𝑝𝑟𝑜𝑗𝑣 𝑢 = 𝑢 − ‖𝑣‖2 𝑣 (vector component of 𝑢 orthogonal to 𝑣).
multiplication defined on 𝑉.
solution, then S is called a linearly independent set. If there are other solutions,
Definition 1.9.4 [9 , 𝑝. 213 ]. ( Span ) The vector 𝑣1 , 𝑣2 , … , 𝑣𝑛 in a vector space 𝑉 are said
vectors are distinct and we denote them as a set 𝑆 = {𝑣1 , 𝑣2 , … , 𝑣𝑛 }, then we also say that
14
Definition 1.9.5 [2 , 𝑝. 259 ]. (row space and column space) If 𝐴 an 𝑚 × 𝑛
matrix, then the subspace of 𝑅𝑛 spanned by the row vectors of 𝐴 is called the row
column space of A.
Definition 1.9.6 [2 , 𝑝. 273 ]. (Rank and nullity) The common dimension of the
row space and column space of a matrix 𝐴 is called the rank of 𝐴 and is denoted
by 𝑟𝑎𝑛𝑘(𝐴), the dimension of the null space which is the solution space of
only if 𝑑𝑒𝑡(𝐴) ≠ 0
15
The matrix 𝐴𝑚×𝑛 = [𝑇(𝑒1 ): 𝑇(𝑒2 ): … : 𝑇(𝑒𝑛 )] is called the standard matrix for the linear
transformation and 𝑇: 𝑅 𝑛 → 𝑅 𝑚 is a multiplication by 𝐴, that is 𝑇𝐴 (𝑥) = 𝐴𝑥.
(a) 𝐴 is invertible
(b) 𝑨𝒙 = 𝟎 has only the trivial solution.
(c) The reduced row-echelon form of 𝐴 is 𝐼𝑛
(d) 𝐴 is expressible as a product of elementary matrices.
(e) 𝑨𝒙 = 𝒃 is consistent for every 𝑛 × 1 matrix 𝑏.
(f) 𝑨𝒙 = 𝒃 has exactly one solution for every 𝑛 × 1 matrix 𝑏.
(g) det(𝐴) ≠ 0
(h) The range of 𝑇𝐴 is 𝑅𝑛
(i) 𝑇𝐴 is one–to–one.
(j) 𝐴 has rank 𝑛.
(k) 𝐴 has nullity 0.
For proofs of these properties [2].
16
Theorem 1.9.15 [9 , 𝑝. 365 ]. Let 𝑇: 𝑅 𝑛 → 𝑅 𝑚 be a linear transformation defined by
Lemma 1.9.16 [10]. Consider the vector spaces 𝑅𝑚 and 𝑅𝑛 , and let 𝑇𝐴 ∶ 𝑅𝑚 → 𝑅𝑛 and
𝑇 𝐴𝑇 : 𝑅 𝑛 → 𝑅 𝑚 the adjoint operator, then the following statement holds
17
Chapter Two
Determinant
There are many ways that general 𝑚 × 𝑛 determinants can be defined. We'll
first define a determinant function in terms of characterizing properties that we
want it to have. Then we'll use the construction of a determinant following the
method given in the section, and through it we will prove that cofactor expansion
definition of the determinant and radic definition are the same.
We generalize the idea given in [8] for the case of rectangular matrices:
That is, 𝐼̃ 𝑚×𝑛 consists of all matrices of the form [𝐴1 𝐴2 ⋯ 𝐴𝑛 ] where
0
𝐴𝑗 ∈ {𝑒1 , 𝑒2 , … , 𝑒𝑚 } ∪ [0], {𝑒1 , 𝑒2 , … , 𝑒𝑚 } is the standard basis for 𝑅 𝑚
⋮
0
Such a matrix is called a permutation matrix in literature.
0 1 0 0 0
Example, 𝐴 = [0 0 1 0 0] is an example of an element in 𝐼̃
0 0 0 0 1
18
where 𝑝 = 1 + 2 + ⋯ + 𝑚 and 𝑞 = 𝑗1 + 𝑗2 + ⋯ + 𝑗𝑛 ( 𝑗1 , 𝑗2 , … , 𝑗𝑛 ) represent columns
A2: If the matrix 𝐵 is identical to the matrix 𝐴 except the entries in one of the
rows of 𝐵 are each equal to the corresponding entries of multiplied by the same
scalar 𝑐, then |𝐵| = 𝑐|𝐴|.
A3: If the matrices 𝐴, 𝐵 and 𝐶 are identical except for the entries in one row, and
for that row an entry in 𝐴 is found by adding the corresponding entries in 𝐵
and 𝐶, then |𝐴| = |𝐵| + |𝐶|.
A4: If the matrix 𝐵 is the result of exchanging two rows of 𝐴, then the determinant
of 𝐵 is the negation of the determinant of 𝐴.
These conditions are enough to characterize the determinant, but they don’t
show such a determinant function exists and is unique. We'll show both existence
and uniqueness, but start with uniqueness. First, we'll note a couple of properties
that determinant functions have that follow from the definition.
(b) The determinant of any matrix 𝐴𝑚×𝑛 ( 𝑚 ≤ 𝑛) with two identical rows is 0.
(c) If one row of a matrix 𝐴𝑚×𝑛 ( 𝑚 ≤ 𝑛) is a scalar multiple of another row, then
its determinant is 0.
(d) If a multiple of one row of a matrix 𝐴𝑚×𝑛 ( 𝑚 ≤ 𝑛) is added to another row,
then the resulting matrix has the same determinant as the original matrix.
Proof: Property (a) follows from the second statement (A2) in the definition.
If 𝐴 has a whole row of 0′ 𝑠, then using that row and 𝑐 = 0 in the second
statement (A2) of the definition, then 𝐵 = 𝐴.
So, |𝐴| = 0|𝐵|. Therefore, det(𝐴) = 0.
Property (b) follows from the fourth statement (A4) in the definition.
19
If you exchange the two identical rows, the result is the original matrix, but its
determinant is negated. The only scalar which is its own negation is 0.
Therefore, the determinant of the matrix is 0.
Property (c) follows from the second statement (A2) in the definition and Property (b).
Property (d) follows from the third statement (A3) in the definition and Property (c).
Proof: The four properties that determinants are enough to find the value of the
determinant of a matrix.
Suppose a matrix 𝐴𝑚×𝑛 (𝑚 ≤ 𝑛) has more than one nonzero entry in a row. Then
using the third statement (A3) in definition 2.1.2.
Now, det(𝐴) = det(𝐴1 ) + det(𝐴2 ) + ⋯ + det(𝐴𝑛 )
Where 𝐴𝑗 is the matrix that looks just like 𝐴 except in that row, all the entries are
0 expect the 𝑗𝑡ℎ one which is the 𝑗𝑡ℎ entry of that row in 𝐴.
That means we can reduce the question of evaluating determinants of general
matrices to evaluating determinants of matrices that have at most one nonzero
entry in each row.
By Property (a) in the theorem 2.1.3, if the matrix has a row of all 0′ 𝑠, its
determinant is 0. Thus, we only need to consider matrices that have exactly one
nonzero entry in each row.
Using the second statement (A2) in definition 2.1.2, we can further assume that
the nonzero entry in that row is 1.
Now, we're down to evaluating determinants that only have entries of 0′ 𝑠 and 1′ 𝑠
with exactly one 1 in each row.
If two of those rows have the 1′ 𝑠 in the same column, then by Property (b) in the
theorem 2.1.3, that matrix has determinant 0.
Now the only matrices left to consider are matrices from 𝐼̃.
Using alternation, the fourth condition (A4) in definition 2.1.2, the rows can be
interchanged until the 1′ 𝑠 only lie on the entry 𝑎𝑖𝑗
20
Finally, we're left with a matrix in 𝐼̃, but by the first condition (A1) in definition
2.1.2, its determinant is (−1)𝑝+𝑞 .
Thus, the value of the determinant of every matrix is determinant by the
definition. There can be only one determinant function.
We need some way to construct a function with those properties, and well do
that with a "cofactor construction" and " radic construction".
The determinant can also be viewed as a function of the columns of the matrix.
Let these columns be 𝐴1 , 𝐴2 , . . , 𝐴𝑛 , then we write the determinant as
|𝐴1 , 𝐴2 , . . , 𝐴𝑛 | or 𝑑𝑒𝑡(𝐴1 , 𝐴2 , . . , 𝐴𝑛 )
= ∑(−1)1+𝑖 𝑎1𝑖
𝑖=1
Example 2.2.2 |1 5 9| = 1 − 5 + 9 = 5
21
Definition 2.2.3 [3] (Minor) Let 𝐴 = (𝑎𝑖𝑗 ) be an 𝑚 × 𝑛 matrix. for each entry
𝑎𝑖𝑗 of 𝐴, we define the minor 𝑀𝑖𝑗 of 𝑎𝑖𝑗 to be the determinant of the
( 𝑚 − 1) × (𝑛 − 1) matrix which remains when the 𝑖𝑡ℎ row and 𝑗𝑡ℎ column are
deleted from 𝐴.
1 3 5
Example 2.2.4 Let 𝐴 = [ ]
2 4 6
To find 𝑀11 , look at element 𝑎11 = 1, delete the entries from column 1 and row 1
that corresponding to 𝑎11 = 1, see the image below.
[
1 − −]
− 4 6
Then 𝑀11 is the determinant of remaining matrix, i.e.,
𝑀11 = |4 6| = 4 − 6 = −2
Similarly, 𝑀22 can be found by looking at the element 𝑎22 = 4 and delete the
same row and column where this element is found, i.e., deleting the second row,
second column:
[
1 − 5]
− 4 −
Then, 𝑀22 = |1 5| = 1 − 5 = −4
It is easy to see that for the matrix 𝐴 the minors of ramming elements are
− 3 −]
[
2 − 6
𝑀12 = |2 6| = 2 − 6 = −4
[− − 5]
2 4 −
𝑀13 = |2 4| = 2 − 4 = −2
− 3 5]
[
2 − −
𝑀21 = |3 5| = 3 − 5 = −2
1 3 −]
[
− − 6
𝑀23 = |1 3| = 1 − 3 = −2
22
Definition 2.2.5 [3] (Cofactor). Let 𝐴 = (𝑎𝑖𝑗 ) be an 𝑚 × 𝑛 matrix. for each
entry 𝑎𝑖𝑗 of 𝐴, we define the cofactor of 𝑎𝑖𝑗 , which is denoted by 𝐶𝑖𝑗 as
The matrix of cofactors of 𝐴 is 𝐶 with 𝑐𝑖𝑗 is cofactor of 𝑎𝑖𝑗 and the adjoint of
𝐴 is 𝑎𝑑𝑗(𝐴) is the transpose of 𝐶.
Basically, the cofactor is either 𝑀𝑖𝑗 , or −𝑀𝑖𝑗 where the sign depends on the
location of the element in the matrix. For that reason, it is easier to know the
pattern of cofactor instead of actually remembering the formula. If you start in the
position corresponding to 𝑎11 with a positive sign, the sign of the cofactor has an
alternating pattern. you can see this by looking at a matrix containing the sign of
the cofactors:
+ − + ⋯ ⋯
[−+ − ⋯ ⋯]
⋮ ⋮ ⋮ ⋮ ⋮
The element 1 in matrix 𝐴 (example 2.2.4) has place sign + and minor -2 so its
cofactor is + (-2) = -2
The element 4 in matrix 𝐴 (example 2.2.4) has place sign + and minor -4 so its
cofactor is + (- 4) = - 4
Proceeding in this way we can find all the cofactors.
The original matrix, its matrix of minors and its matrix of cofactors are:
1 3 5 −2 −4 −2
𝐴= [ ], 𝑀=[ ],
2 4 6 −2 −4 −2
−2 2
−2 4 −2
𝐶= [ ], 𝑎𝑑𝑗(𝐴) = [ 4 −4]
2 −4 2
−2 2
23
Definition 2.3.1 [3] Let 𝐴 = (𝑎𝑖𝑗 ) be an 𝑚 × 𝑛 matrix with 𝑚 ≤ 𝑛. (horizontal
matrix). The determinant of 𝐴 is defined as
𝑛
𝑑𝑒𝑡(𝐴) = 𝑎11 𝐶11 + 𝑎12 𝐶12 + 𝑎13 𝐶13 + ⋯ + 𝑎1𝑛 𝐶1𝑛 = ∑ 𝑎1𝑗 𝐶1𝑗
𝑗=1
horizontal matrix 𝐴𝑇 .
The following theorem asserts that we can evaluate the determinant of a larger
horizontal matrix by selecting any row, multiplying each element in that row by
its corresponding cofactor, and summing the result, a result which is true in the
case of square matrices (Theorem 1.4.2).
The following theorem is from [7] but we give here another proof
= ∑ 𝑎𝑖𝑗 𝐶𝑖𝑗
𝑗=1
This is called the determinant using cofactor expansion along the 𝑖𝑡ℎ row.
The proof of this theorem will be given after proving Theorem 2.4.1
1 3 5
Example 2.3.3. Evaluate the determinant of 𝐴 = [ ] by cofactor expansion
2 4 6
1 3 5
𝑑𝑒𝑡(𝐴) = | | = 1 |4 6| − 3 |2 6| + 5 |2 4|
2 4 6
24
= 1 ( −2 ) − 3 (−4) + 5(−2 ) = 0
1 3 5
𝑑𝑒𝑡(𝐴) = | | = −2 |3 5| + 4 |1 5| − 6 |1 3|
2 4 6
= −2 ( −2 ) + 4 (−4) − 6(−2 ) = 0
2 3 4 1
Example 2.3.4 Evaluate the determinant of 𝐴 = [1 5 0 2] using Theorem 2.3.2
2 4 1 3
Solution: Since 𝑚 < 𝑛, expanding a long any row, say first row
2 3 4 1
5 0 2| − 3 |1 0 2| + 4 |1 5 2| − 1 |1 5 0|
|1 5 0 2| = 2 |
4 1 3 2 1 3 2 4 3 2 4 1
2 4 1 3
det(𝐴) = 0
25
2.4 Effect of elementary row operations on determinates
definition 2.3.1 involves the summation of 𝑛𝑝( 𝑚−1) , each term being a product
so another technique has been devised to evaluate the determinant which works
quite efficiently. This technique uses row operations to put a matrix into a form in
which the determinant is easily calculated, keeping track of the row operations
used, and how they affect the determinant, we can backtrack, and determine what
The following theorem is from [3], [7], but we give here another proof
Theorem 2.4.1. If 𝐴 and 𝐵 are 𝑚 × 𝑛 matrices with 𝑚 ≤ 𝑛, and 𝐵 is obtained
(−1)𝑛 𝑎1𝑛 ) + 𝑎23 (𝑎11 − 𝑎12 + ⋯ + (−1)𝑛 𝑎1𝑛 ) + … + (−1)𝑛+1 𝑎2𝑛 (𝑎11 −
=(𝑎21 𝑎12 − 𝑎21 𝑎13 + ⋯ + (−1)𝑛 𝑎21 𝑎1𝑛 ) − (𝑎22 𝑎11 − 𝑎22 𝑎13 +
⋯ + (−1)𝑛 𝑎22 𝑎1𝑛 ) + (𝑎23 𝑎11 − 𝑎23 𝑎12 + ⋯ + (−1)𝑛 𝑎23 𝑎1𝑛 ) + … +
26
Also ,
det(𝐴) = 𝑎11 (𝑎22 − 𝑎23 + ⋯ + (−1)𝑛 𝑎2𝑛 ) − 𝑎12 (𝑎21 − 𝑎23 + ⋯ +
(−1)𝑛 𝑎2𝑛 ) + 𝑎13 (𝑎21 − 𝑎22 + ⋯ + (−1)𝑛 𝑎2𝑛 ) + … + (−1)1+𝑛 𝑎1𝑛 (𝑎21 −
= (𝑎11 𝑎22 − 𝑎11 𝑎23 + ⋯ + (−1)𝑛 𝑎11 𝑎2𝑛 ) − (𝑎12 𝑎21 − 𝑎12 𝑎23 +
𝑛
⋯ + (−1)𝑛 𝑎12 𝑎2𝑛 ) + (𝑎13 𝑎21 − 𝑎13 𝑎22 + ⋯ + (−1) 𝑎13 𝑎2𝑛 ) + … +
a matrix obtained from 𝐴 by swapping two rows. Say row 𝑟 and 𝑟 + 1 of 𝐴 were
We may evaluate 𝑑𝑒𝑡 (𝐵) by cofactor expansion along its first row.
To compute 𝑑𝑒𝑡(𝐵), we will need to look at the submatrices 𝐵(𝑖 , 𝑗 ). Our choice
of 𝑖 = 1 means that 𝐵(1 , 𝑗 ) can be obtained from 𝐴(1 , 𝑗 ) by swapping the rows
matrix that is obtained from 𝐴(1 , 𝑗 ) by swapping two rows, and thus, by our
27
Now,
𝑑𝑒𝑡 (𝐵) = 𝑏11 (−1)1+1 𝑑𝑒𝑡 𝐵(1 , 1 ) + ⋯ + 𝑏1𝑛 (−1)1+𝑛 𝑑𝑒𝑡 𝐵(1 , 𝑛 )
det(𝐵) = − det(𝐴)
This proves the result for the interchange of two adjacent rows in an 𝑚 × 𝑛
matrix. To see that this result holds for arbitrary row interchanges, we note that
the interchange of two rows, say row 𝑟 and 𝑠 where 𝑟 < 𝑠 can be performed by
and each one changes the sign of the determinant, the net effect is a change of
sign as desired.
Proof: Let 𝐵 be the matrix obtained by moving the 𝑖𝑡ℎ row of 𝐴 to the top, using
𝑖−1 interchanges of adjacent rows. Thus 𝑑𝑒𝑡 (𝐵) = (−1)𝑖−1 det(𝐴), but
𝑏1𝑗 = 𝑎𝑖𝑗 and 𝐵1𝑗 = 𝐴𝑖𝑗 for 𝑗 ∈ [1, 𝑛] and so
Hence,
𝑛
det(𝐴) = (−1)𝑖−1 det(𝐵) = (−1) ∑(−1)1+𝑗 𝑏
𝑖−1
1𝑗 det(𝐵1𝑗 )
𝑗=1
28
𝑛 𝑛
Giving the formula for cofactor expansion along the 𝑖𝑡ℎ row.
Corollary 2.4.2 [3] If any two rows of a horizontal matrix are identical, then the
value of its determinant is zero.
Proof: Let |𝐴| be the determinant of the horizontal matrix 𝐴. Assume that row 𝑖
and row 𝑗 in 𝐴 are identical. By Theorem 2.4.1 interchange row 𝑖 and row 𝑗, the
determinant of the resulting matrix is −|𝐴|. But the original matrix and the
resulting matrix are the same
Proof: If we expand along the 𝑖𝑡ℎ row of 𝐵 to calculate its determinant, we get
But the reason we have chosen the 𝑖𝑡ℎ row of 𝐵 is that we know that 𝑏𝑖𝑗 = 𝑘 𝑎𝑖𝑗
for 𝑗 = 1 , … , 𝑛. Moreover, since the submatrices 𝐵(𝑖 , 𝑗 ) will all have row 𝑖
removed, and since this is the only place where 𝐵 differs from 𝐴, we see that
𝐴(𝑖 , 𝑗 ) = 𝐵(𝑖 , 𝑗 ). Thus, the cofactor 𝐵𝑖𝑗 for 𝑏𝑖𝑗 is the same as the cofactor 𝐴𝑖𝑗
29
𝑖+𝑛
det(𝐵) = 𝑘𝑎𝑖1 𝐵𝑖1 + ⋯ + (−1) 𝑘 𝑎𝑖𝑛 𝐵𝑖𝑛
= 𝑘 (𝑎𝑖1 𝐴𝑖1 + ⋯ + (−1)𝑖+𝑛 𝑎𝑖𝑛 𝐴𝑖𝑛 )
= 𝑘 det(𝐴)
If we know the determinant of matrix 𝐴, we can use this information to calculate the
det(𝑘 𝐴) = 𝑘𝑚 det(𝐴)
Proof: Since all 𝑚 rows of 𝐴 are multiplied by the scalar 𝑘 to get 𝑘 𝐴, using the
above theorem 𝑚 times gives
= 𝑘 𝑚 det(𝐴)
30
Then we can compute the determinant of 𝐵 by expanding along row 𝑟, getting
Our choice of 𝑟 gets us that 𝑏𝑟𝑗 = 𝑘𝑎𝑠𝑗 + 𝑎𝑟𝑗 for all 𝑗 = 1, … , 𝑛. So, now let's
consider the submatrices 𝐵(𝑟 , 𝑗 ) and 𝐴(𝑟 , 𝑗 ) are the cofactor of 𝐴 and 𝐵. This
means that 𝐵(𝑟 , 𝑗 ) can obtained from 𝐴(𝑟 , 𝑗 ) by adding 𝑘 times row 𝑠 to row 𝑟.
31
Since we know how elementary row operations affect the determinant, we can
compute the determinant of a given matrix by computing the determinant of its
r.r.e.f (see definition 1.7.1) and taking into account the effect of the row
operations. The same procedure that we used in books of linear algebra for
determinant of square matrices.
The following table describes the effect of applying row operations on
computing the determinant of a horizontal matrix.
We mention here that these properties correspond to their counter parts for
determinants of square matrices (See Theorem 1.3.5)
2 4 6
Example 2.4.6 Find the determinant of 𝐴 = [ ]
1 3 5
Solution: We use row reduction until 𝐴 is in reduced row echelon form. At each
2 4 6 𝑅1 ↔𝑅2 : det×(−1) 1 3 5
| | → −| |
1 3 5 2 4 6
1 1
R2 −2𝑅1 →R2 ∶det 𝑢𝑛𝑐ℎ𝑎𝑛𝑔𝑒𝑑 (− )𝑅2 : 𝑑𝑒𝑡(− )
− |1 3 5 1 3 5
2 2
→ | → −| |
0 −2 −4 0 1 2
32
Note that, the definition of the determinant (see theorem 2.4.1, 2.4.3, 3.1.1) satisfies the
Also, the cofactor definition of the determinant (see theorem 2.3.5, 2.4.2, 2.4.5) satisfies
𝑎1𝑗1 ⋯ 𝑎1𝑗𝑚
𝑟+𝑆
det(𝐴) = ∑ (−1) 𝑑𝑒𝑡 [ ⋮ ⋱ ⋮ ] (2.5.1)
1≤𝑗1 <⋯<𝑗𝑚 ≤𝑛 𝑎𝑚𝑗1 ⋯ 𝑎𝑚𝑗𝑚
Where 𝑗1 , 𝑗2 , … , 𝑗𝑚 ∈ 𝑁, 𝑟 = 1 + 2 + ⋯ + 𝑚 and 𝑠 = 𝑗1 + 𝑗2 + ⋯ + 𝑗𝑚 .
33
Proof: Let 𝐴 be an 𝑚 × 𝑛 matrix with 𝑚 ≤ 𝑛, by Radic's definition
𝑎1𝑗1 ⋯ 𝑎1𝑗𝑚
det(𝐴) = ∑ (−1)𝑟+𝑆 𝑑𝑒𝑡 [ ⋮ ⋱ ⋮ ]
1≤𝑗1 <⋯<𝑗𝑚 ≤𝑛 𝑎𝑚𝑗1 ⋯ 𝑎𝑚𝑗𝑚
Where 𝑗1 , 𝑗2 , … , 𝑗𝑚 ∈ 𝑁, 𝑟 = 1 + 2 + ⋯ + 𝑚 and 𝑠 = 𝑗1 + 𝑗2 + ⋯ + 𝑗𝑚 .
(1) If a row of matrix 𝐴 is a linear combination of some other rows, then all
𝑎1𝑗1 ⋯ 𝑎1𝑗𝑚
𝑑𝑒𝑡 [ ⋮ ⋱ ⋮ ] contains a row is a linear combination of some other rows, and
𝑎𝑚𝑗1 ⋯ 𝑎𝑚𝑗𝑚
𝑎1𝑗1 ⋯ 𝑎1𝑗𝑚
therefore, all 𝑑𝑒𝑡 [ ⋮ ⋱ ⋮ ] = 0, where 𝑗1 , 𝑗2 , … , 𝑗𝑚 ∈ 𝑁, are square matrices,
𝑎𝑚𝑗1 ⋯ 𝑎𝑚𝑗𝑚
hence 𝑑𝑒𝑡(𝐴) = 0
𝑎11 𝑎12
⋯ 𝑎1𝑛
⋮ ⋮
⋱ ⋮
𝑎
(2) Let 𝐴 = 𝑖1 𝑎𝑖2
⋯ 𝑎𝑖𝑛 , and suppose that 𝐵 is the matrix obtained
⋮ ⋮
⋱ ⋮
[𝑎𝑚1 𝑎𝑚2
⋯ 𝑎𝑚𝑛 ]
𝑎11 𝑎12 ⋯ 𝑎1𝑛
⋮ ⋮ ⋱ ⋮
from 𝐴 by multiplying row 𝑖 by 𝑘 , 𝐵 = 𝑘𝑎𝑖1 𝑘𝑎𝑖2 ⋯ 𝑘𝑎𝑖𝑛 then,
⋮ ⋮ ⋱ ⋮
[ 𝑎𝑚1 𝑎𝑚2 ⋯ 𝑎𝑚𝑛 ]
34
𝑎1𝑗1 𝑎1𝑗2 ⋯ 𝑎1𝑗𝑚 𝑎1𝑗1 𝑎1𝑗2 ⋯ 𝑎1𝑗𝑚
⋮ ⋮ ⋱ ⋮ ⋮ ⋮ ⋱ ⋮
𝑑𝑒𝑡 𝑘𝑎 𝑖𝑗1 𝑘𝑎𝑖𝑗2 ⋯ 𝑘𝑎𝑖𝑗𝑚 = 𝑘𝑑𝑒𝑡 𝑎𝑖𝑗1 𝑎𝑖𝑗2 ⋯ 𝑎𝑖𝑗𝑚 , 𝑗1 , 𝑗2 , … , 𝑗𝑚 ∈ 𝑁
⋮ ⋮ ⋱ ⋮ ⋮ ⋮ ⋱ ⋮
[ 𝑎𝑚𝑗1 𝑎𝑚𝑗2 ⋯ 𝑎𝑚𝑗𝑚 ] 𝑎
[ 𝑚𝑗1 𝑎𝑚𝑗2 ⋯ 𝑎𝑚𝑗𝑚 ]
can be taken 𝑘 out from all determinant square in common to produce 𝑑𝑒𝑡(𝐵) = 𝑘𝑑𝑒𝑡(𝐴).
then,
𝑎1𝑗1 𝑎1𝑗2 ⋯ 𝑎1𝑗𝑚
⋮ ⋮ ⋱ ⋮
𝑎𝑠𝑗1 𝑎𝑠𝑗2 𝑎𝑠𝑗𝑚
⋯
det(𝐵) = ∑ (−1)𝑟+𝑆 𝑑𝑒𝑡 ⋮ ⋮ ⋱ ⋮ ,
𝑎𝑟𝑗1 𝑎𝑟𝑗2 ⋯ 𝑎𝑟𝑗𝑚
1≤𝑗1 <⋯<𝑗𝑚 ≤𝑛
⋮ ⋮ ⋱ ⋮
𝑎 𝑎 ⋯ 𝑎𝑚𝑗𝑚 ]
[ 𝑚𝑗1 𝑚𝑗2
35
can be taken sign negative out from all determinant square in common to produce
𝑑𝑒𝑡(𝐵) = −𝑑𝑒𝑡(𝐴).
(5) We need to show that radic definition of the determinant satisfies the axioms
(A3) of determinant function
𝑎11 𝑎12 ⋯ 𝑎1𝑛 𝑏11 𝑏12 ⋯ 𝑏1𝑛
𝑎21 𝑎22 ⋯ 𝑎2𝑛 𝑎21 𝑎22 ⋯ 𝑎2𝑛
Let 𝐴 = [ ⋮ ⋱ ⋮ ],
⋮ 𝐵=[ ],
⋮ ⋮ ⋱ ⋮
𝑎𝑚1 𝑎𝑚2 ⋯ 𝑎𝑚𝑛 𝑎𝑚1 𝑎𝑚2 ⋯ 𝑎𝑚𝑛
𝑎11 + 𝑏11 𝑎12 + 𝑏12 ⋯ 𝑎1𝑛 + 𝑏1𝑛
𝑎21 𝑎22 ⋯ 𝑎2𝑛
𝐶=[ ]
⋮ ⋮ ⋱ ⋮
𝑎𝑚1 𝑎𝑚2 ⋯ 𝑎𝑚𝑛
36
Then,
𝑎1𝑗1 + 𝑏 𝑎1𝑗2 + 𝑏 ⋯ 𝑎1𝑗 + 𝑏1𝑗
1𝑗1 1𝑗2 𝑚 𝑚
𝑎2𝑗1 𝑎2𝑗2 ⋯ 𝑎2𝑗𝑚
det(𝐶) = ∑ (−1)𝑟+𝑆 𝑑𝑒𝑡
⋮ ⋮ ⋱ ⋮
1≤𝑗1 <⋯<𝑗𝑚 ≤𝑛
[ 𝑎𝑚𝑗1 𝑎𝑚𝑗2 ⋯ 𝑎 𝑚𝑗𝑚 ]
Hence, det(𝐶) =
𝑎1𝑗1 𝑎1𝑗2 ⋯ 𝑎1𝑗𝑚 𝑏1𝑗1 𝑏1𝑗2 ⋯ 𝑏1𝑗𝑚
𝑎2𝑗1 𝑎2𝑗2 ⋯ 𝑎2𝑗𝑚 𝑎2𝑗1 𝑎2𝑗2 ⋯ 𝑎2𝑗𝑚
∑ (−1)𝑟+𝑆 ( 𝑑𝑒𝑡 + 𝑑𝑒𝑡 )
⋮ ⋮ ⋱ ⋮ ⋮ ⋮ ⋱ ⋮
1≤𝑗1 <⋯<𝑗𝑚 ≤𝑛
[𝑎𝑚𝑗1 𝑎𝑚𝑗2 ⋯ 𝑎𝑚𝑗𝑚 ] [𝑎𝑚𝑗1 𝑎𝑚𝑗2 ⋯ 𝑎𝑚𝑗𝑚 ]
= det(𝐴) + det(𝐵)
Note that, the definition of the determinant (see properties 2,3,5) satisfies the
axioms of determinant function 2.1.2 (**).
Also, the Radic definition of the determinant (see properties 1.4 ) satisfies the
37
Corollary 2.5.2 The determinant obtained by cofactor expansion and Radic
definition are the same.
Proof: the cofactor definition and Radic definition are determinant function, since
the four characterizing properties of determinant listed in definition 2.1.2 are
satisfied by the cofactor definition 2.3.1 and Radic definition 2.5.1 of
determinants, and because of uniqueness of determinant function (see theorem
2.1.4), the cofactor definition and Radic definition are the same. ( see * , **)
= 𝑎1 − 𝑎2 + 𝑎3
𝑎1 𝑎2 𝑎3
Example 2.5.4 Evaluate the determinant of 𝐴 = [𝑏 𝑏2 𝑏3 ] using Radic definition
1
Solution :
𝑎1 𝑎2 𝑎1 𝑎3 𝑎2 𝑎3
|𝐴| = (−1)(3)+(1+2) |𝑏 | (−1) (3)+(1+3)
| | (−1) (3)+(2+3)
|
1 𝑏2 + 𝑏1 𝑏3 + 𝑏2 𝑏3 |
𝑎1 𝑎2 𝑎1 𝑎3 𝑎2 𝑎3
= |𝑏 𝑏2 | − | 𝑏1 𝑏3 | + | 𝑏2 𝑏3 |
1
38
Theorem 2.5.5 [13 ] Let 𝐴 = [𝐴1 , … , 𝐴𝑛 ] be 2 × 𝑛 matrix with 𝑛 ≥ 2. Then
𝑑𝑒𝑡(𝐴2 , 𝐴3 − 𝐴4 + … + (−1)𝑛−1 𝐴𝑛 ) + ⋯ +
𝑑𝑒𝑡(𝐴𝑛−1 , 𝐴𝑛 )
Base case, 𝑛 = 3,
39
Theorem 2.5.5 converts the computations of the determinant of an 2 × 𝑛
matrix according to Radic definition which needs computing (𝑛2) determinants to
computation of 𝑛 − 1 determinants of size 2 × 2.
𝑛(𝑛−1)
( (𝑛2) = > 𝑛−1 )
2
1 3 5
Example 2.5.6 Evaluate the determinant of 𝐴= [ ] using theorem 2.5.5
2 4 6
1 −2 3 5
Then, 𝑑𝑒𝑡(𝐴) = | |+ | | = (−2 + 4 ) + ( 18 − 20 ) = 0 .
2 −2 4 6
Base case: 𝑛 = 3,
= 𝑑𝑒𝑡(𝐴1 , 𝐴2 ) − 𝑑𝑒𝑡(𝐴1 − 𝐴2 , 𝐴3 )
𝑑𝑒𝑡(𝐴1 , … , 𝐴𝑘 )
40
We will show that the equality holds for 𝑛 = 𝑘 + 1
+ 𝑑𝑒𝑡(𝐴𝑘 , 𝐴𝑘+1 )
1 3 5
Example 2.5.8. Evaluate the determinant of 𝐴 = [ ], using theorem 2.5.7
2 4 6
Solution :
41
Chapter Three
Theorem 3.1.1 If every element in any fixed row of a horizontal matrix can be
expressed as the sum of tow quantities then the given horizontal matrix
determinant can be expressed as the sum of tow horizontal matrix determinant of
the same order with the elements of the remaining rows of the both being the
same.
42
𝑎22 … 𝑎2𝑛 𝑎21 … 𝑎2𝑛
| 𝐴 | = (𝛼 + 𝑎11 ) | ⋮ ⋱ ⋮ | − (𝛽 + 𝑎12 ) | ⋮ ⋱ ⋮ |+
𝑎𝑚2 … 𝑎𝑚𝑛 𝑎𝑚1 … 𝑎𝑚𝑛
𝑎21 … 𝑎2𝑛 𝑎21 … 𝑎2(𝑛−1)
1+𝑛
(𝛾 + 𝑎13 ) | ⋮ ⋱ ⋮ | + ⋯ +(−1) (𝛿 + 𝑎1𝑛 ) | ⋮ ⋱ ⋮ |
𝑎𝑚1 … 𝑎𝑚𝑛 𝑎𝑚1 … 𝑎𝑚(𝑛−1)
det(𝐴) =
𝑎22 … 𝑎2𝑛 𝑎21 … 𝑎2𝑛 𝑎21 … 𝑎2𝑛
(𝛼 | ⋮ ⋱ ⋮ |− 𝛽 | ⋮ ⋱ ⋮ |+ 𝛾| ⋮ ⋱ ⋮ |+
𝑎𝑚2 … 𝑎𝑚𝑛 𝑎𝑚1 … 𝑎𝑚𝑛 𝑎𝑚1 … 𝑎𝑚𝑛
𝑎21 … 𝑎2(𝑛−1) 𝑎22 … 𝑎2𝑛 𝑎21 … 𝑎2𝑛
⋯ +(−1)1+𝑛 𝛿| ⋮ ⋱ ⋮ |) + (𝑎11 | ⋮ ⋱ ⋮ | − 𝑎12 | ⋮ ⋱ ⋮ |+
𝑎𝑚1 … 𝑎𝑚(𝑛−1) 𝑎𝑚2 … 𝑎𝑚𝑛 𝑎𝑚1 … 𝑎𝑚𝑛
2- This property is valid for square matrices as well ( see theorem 1.3.6)
By Radic definition
Where 𝑟 = 1 + 2 + ⋯ + 𝑚 and 𝑠 = 𝑗1 + 𝑗2 + ⋯ + 𝑗𝑚
43
𝑟+𝑆
det(𝐴 𝐵) = ∑ (−1) det(A) 𝑑𝑒𝑡 ([Bj1 , … , Bj𝑚 ])
1≤𝑗1 <⋯<𝑗𝑚 ≤𝑛
= det(A)det(B)
2 2 3 5 7
Example 3.1.3 Prove theorem 3.1.2 for 𝐴 = [ ], 𝐵=[ ]
−1 3 2 1 4
10 12 22
Then, 𝐴 𝐵 = [ ],
3 −2 5
In fact there is no determinant function that satisfies 𝑑𝑒𝑡(𝐴𝐵) = 𝑑𝑒𝑡(𝐴)𝑑𝑒𝑡(𝐵) for all
matrices 𝐴, 𝐵.
1 2 3
1 0 2
Example 3.1.4 Let 𝐴=[ ], 𝐵 = [3 0 −2]
−1 1 1
2 1 0
5 3 3
Then, 𝐴𝐵= [ ],
4 0 −5
We notice from this example that the determinant is not distributed in the case
that the first matrix is rectangular, so for the theorem to be true, the first matrix
must be the square.
44
0 1
1 1 0
Example 3.1.5 Let 𝐴=[ ], 𝐵 = [ 2 0]
2 3 0
−4 2
2 3 0
2 1
Then, 𝐴 𝐵 = [ ], 𝐵 𝐴 = [2 2 0 ]
6 2
0 2 0
−2 = (1)(−2) = (−2)(1) ≠ 0
𝑑𝑒𝑡
𝑚 + 𝑛 odd, then 1≤𝑖≤𝑚 [(𝑎𝑖𝑗 )] = 0, where 𝑎1𝑗 = 1 for all 𝑗, 1 ≤ 𝑗 ≤ 𝑛 .
1≤𝑗≤𝑛
1 ≤ 𝑚 < 𝑛.
1 ≤ 𝑚 < 𝑛,
𝑎2,𝑗
1 1 … 1
𝐷 = 𝑑𝑒𝑡[𝐴1 , 𝐴2 , … , 𝐴𝑛 ] = 𝑑𝑒𝑡 [ ] where 𝐵𝑗 = [ ⋮ ] , (1 ≤ 𝑗 ≤ 𝑛)
𝐵1 𝐵2 … 𝐵𝑛 𝑎𝑚,𝑗
Induction step : We will show that the identity holds for 𝑛+2
1 1 … 1 1 1
𝐷 = 𝑑𝑒𝑡[𝐴1 , 𝐴2 , … , 𝐴𝑛 , 𝐴𝑛+1 , 𝐴𝑛+2 ] = 𝑑𝑒𝑡 [ ]
𝐵1 𝐵2 … 𝐵𝑛 𝐵𝑛+1 𝐵𝑛+2
45
𝐷 = 𝑑𝑒𝑡[𝐵2 , 𝐵3 , … , 𝐵𝑛+2 ] − 𝑑𝑒𝑡[𝐵1 , 𝐵3 , … , 𝐵𝑛+2 ] + ⋯
By theorem 2.5.7, all the resulting determinants are deleted from each other.
Then , 𝐷 = 0.
Another case can be established in the same way(odd 𝑛 and even 𝑚).
This property does not apply if the matrices are square ( 𝑛 × 𝑛) because the
sum of the order is even (whether 𝑛 is odd or even)
1 1 1 1 1
Example 3.1.7 Evaluate the determinant of 𝐴 = [ ] and 𝐵 = [ ]
2 3 4 5 6
Lemma 3.1.8 [1 ] .
𝑇
Where 𝑚 ≤ 𝑛, 𝐴𝐾 = [𝑎1,𝑘 , … , 𝑎𝑚,𝑘 ] for 𝑘 ∈ {1 , … , 𝑛} − {𝑗} and 0𝑚 is an 𝑚
by 1 zero vector .
𝑎1,𝑗
𝑎2,𝑗
𝑎2,𝑗 𝑎1,𝑗
Let 𝐴𝑗 = [ ⋮ ] , 𝐴𝑗 = [ 𝐵 ] where 𝐵𝑗 = [ ⋮ ] , (1 ≤ 𝑗 ≤ 𝑛)
𝑗 𝑎𝑚,𝑗
𝑎𝑚,𝑗
𝑎 𝑎 ⋯ 𝑎1𝑛 0
D = det [𝐴1 , 𝐴2 , … , 𝐴𝑛 , 0𝑚 ] = 𝑑𝑒𝑡 [ 𝐵11 𝐵12 ⋯ 𝐵 ]
1 2 𝑛 0𝑚
46
𝐷 = (−1)2 𝑎1,1 𝑑𝑒𝑡[𝐵2 , 𝐵3 , … , 𝐵𝑛 , 0𝑚 ] + (−1)3 𝑎12 𝑑𝑒𝑡[𝐵1 , 𝐵3 , … , 𝐵𝑛 , 0𝑚 ] + ⋯
Induction hypothesis: Assume that for all 𝑛 and 𝑚, 1 ≤ 𝑚 < 𝑛, it is true that
Induction step: We will show that the identity holds for 𝑛 + 1, 1 ≤ 𝑚 < 𝑛 + 1 ?
𝑎1,𝑗
𝑎2,𝑗
𝑎2,𝑗 𝑎1,𝑗
Let 𝐴𝑗 = [ ⋮ ] , 𝐴𝑗 = [ 𝐵 ] where 𝐵𝑗 = [ ⋮ ] , (1 ≤ 𝑗 ≤ 𝑛)
𝑗 𝑎𝑚,𝑗
𝑎𝑚,𝑗
47
By inductive hypothesis
This property does not apply if the matrices are square since if all the elements
of a column are zeros, then the value of the determinant is zero. (see theorem
1.3.5)
3 5 7
Let 𝐴 = [ ], det(𝐴) = 8
2 1 4
So, |3 5 7 0| = |3 5 7| = 8,
2 1 4 0 2 1 4
and |3 0 5 7| = |3 −5 −7 | = 8
2 0 1 4 2 −1 −4
det[𝐴1 + 𝑋 , … , 𝐴𝑛 + 𝑋] = 𝑑𝑒𝑡[𝐴1 , … , 𝐴𝑛 ]
Proof: by (P.M.I )
48
Induction hypothesis: Assume the assertion is true for all 𝑛 even and 𝑚 odd with
1≤𝑚<𝑛,
det[𝐴1 + 𝑋 , … , 𝐴𝑛 + 𝑋] = 𝑑𝑒𝑡[𝐴1 , … , 𝐴𝑛 ]
With 1 ≤ 𝑚 < 𝑛 + 2 ,
𝑎2𝑗 𝑥2
𝑎1𝑗 𝑥1
Let 𝐴𝑗 = [ 𝐵 ] where 𝐵𝑗 = [ ⋮ ] , (1 ≤ 𝑗 ≤ 𝑛 + 2) and 𝑋 = [ / ] where 𝑋/ = [ ⋮ ]
𝑗 𝑎𝑚𝑗 𝑋 𝑥𝑚
𝑎1,1 + 𝑥1 … 𝑎1,𝑛+2 + 𝑥1
det[𝐴1 + 𝑋 , … , 𝐴𝑛+2 + 𝑋] = 𝑑𝑒𝑡 [ ]
𝐵1 + 𝑋/ … 𝐵𝑛+2 + 𝑋/
𝑎1,1 … 𝑎1,𝑛+2 1 … 1
𝑑𝑒𝑡[𝐴1 + 𝑋 , … , 𝐴𝑛+2 + 𝑋] = 𝑑𝑒𝑡 [ 𝐵 … 𝐵𝑛+2 ] + 𝑥1 𝑑𝑒𝑡 [ ]
1 𝐵1 … 𝐵𝑛+2
= 𝑑𝑒𝑡[𝐴1 , … , 𝐴𝑛+2 ]
49
3 5 7 1
Example 3.1.11 Let 𝐴= [ ], take 𝑋 = [ ],
2 1 4 2
Also, det(𝐴1 + 𝑋 , 𝐴2 + 𝑋 , 𝐴3 + 𝑋) = |4 6 8 | = 8
4 3 6
1 ≤ k ≤ n , we have
3 5 7
Example 3.1.13 let 𝐴 = [ ] , then det(𝐴 ) = 8
2 1 4
det(𝐴 ) = | 2 2| = 8
−1 3
Note, This properties (theorem 3.1.10 and corollary 3.1.12) do not apply if the
matrices are square ( 𝑛 × 𝑛) because the sum of the order is even (Whether 𝑛 is
odd or even )
50
3.2 How determinant is affected by operations on columns.
Where 𝑟 = 1 + 2 + ⋯ + 𝑚
We separate the sum of determinants into tow sums: the first one consisting of the
determinants of matrices which contain the column 𝐴𝑘 = 𝐵𝑘 + 𝐶𝑘 and the
second one consisting of other determinants.
51
= ∑ (−1)𝑟+𝑗1 +𝑗2 +⋯+𝑗𝑚 |𝐴𝑗1 , … , 𝐵𝑘 , … , 𝐴𝑗𝑚 |
1≤𝑗1 <⋯<𝑗𝑚 ≤𝑛
𝑘∈ {𝑗1 ,… ,𝑗𝑚 }
Now the third sum is added and subtracted so that it can be included into both the
first and the second sum :
|𝐵1 + 𝐶1 , 𝐴2 , 𝐴3 |
(2) Interchanging columns in a square matrix results in changing the sign of the
52
Theorem 3.2.3 [12 ]. Let 𝐴 = [𝐴1 , 𝐴2 , … , 𝐴𝑚 , 𝐴𝑚+1 ] be a 𝑚 × (𝑚 + 1)
(𝑚+1)(𝑚+2)
|𝐴| = (−1)(𝑟+ 2
−𝑖)
|𝐴1 , 𝐴2 , … , 𝐴𝑖−1 , 𝐴𝑖+1 , … , 𝐴𝑗−1 , 𝐴𝑗 , 𝐴𝑗+1 , … , 𝐴𝑚+1 |
(𝑚+1)(𝑚+2)
+ (−1)(𝑟+ 2
−𝑗)
|𝐴1 , 𝐴2 , … , 𝐴𝑖−1 , 𝐴𝑖 , 𝐴𝑖+1 , … , 𝐴𝑗−1 , 𝐴𝑗+1 , … , 𝐴𝑚+1 |
Notice that exactly 𝑗 − 𝑖 − 1 inversions are needed to move the column 𝐴𝑗 to the
position between 𝐴𝑖−1 and 𝐴𝑖+1 in the first summand. Similarly, in the second
summand, also 𝑗 − 𝑖 − 1 inversions are needed to move the column 𝐴𝑖 to the
position between 𝐴𝑗−1 and 𝐴𝑗+1 .
In other summands we can simply interchange columns 𝐴𝑖 and 𝐴𝑗 with the sign
change ( square matrix 𝑚 × 𝑚). Thus, we have
(𝑚+1)(𝑚+2)
(𝑟+ −𝑖+(𝑖−𝑗−1))
|𝐴| = (−1) 2
53
𝑟+𝑗1 +𝑗2 +⋯+𝑗𝑚
− ∑ (−1)(−1) |𝐴𝑗1 , … , 𝐴𝑗 , … , 𝐴𝑖 , . . , 𝐴𝑗𝑚 |
1≤𝑗1 <⋯<𝑗𝑚 ≤𝑛
𝑖 ,𝑗 ∈ {𝑗1 ,… ,𝑗𝑚 }
(𝑚+1)(𝑚+2)
det(𝐴) = −(−1)(𝑟+ 2 −𝑗)
|𝐴1 , 𝐴2 , … , 𝐴𝑖−1 , 𝐴𝑗 , 𝐴𝑖+1 , … , 𝐴𝑗−1 , 𝐴𝑗+1 , … , 𝐴𝑚+1 |
(𝑚+1)(𝑚+2)
− (−1)(𝑟+ 2 −𝑖)
|𝐴1 , 𝐴2 , … , 𝐴𝑖−1 , 𝐴𝑖+1 , … , 𝐴𝑗−1 , 𝐴𝑖 , 𝐴𝑗+1 , … , 𝐴𝑚+1 |
𝑛 − 𝑚 > 1 the sum det(𝐴) + det (𝐴/ ) is not zero as explained in the following
example.
1 2 7 4 3
Example 3.2.5 Let 𝐴 = [ ] , det(𝐴) = 5
3 0 1 −1 2
Theorem 3.2.6 [1] (Cyclic). If 1 ≤ 𝑚 < 𝑛, and 𝑚 + 𝑛 is an odd integer, then for all
𝑖 ∈ {1 , … , 𝑛}. We have
54
Applying Theorem 3.1.10 with 𝑋 = −𝐴𝑛 and Lemma 3.1.8 we have
= 𝑑𝑒𝑡[ 𝐴1 − 𝐴𝑛 , … , 𝐴𝑛−1 − 𝐴𝑛 ]
= 𝑑𝑒𝑡[ 𝐴1 − 𝐴𝑛 , … , 𝐴𝑛−1 − 𝐴𝑛 , 𝐴𝑛 − 𝐴𝑛 ]
= 𝑑𝑒𝑡[𝐴𝑛 − 𝐴𝑛 , − 𝐴1 − 𝐴𝑛 , … , − 𝐴𝑛−1 − 𝐴𝑛 , 0𝑚 − 𝐴𝑛 ]
= 𝑑𝑒𝑡[ 𝐴1 + 𝐴𝑛 , … , 𝐴𝑛−1 + 𝐴𝑛 , 𝐴𝑛 ]
= 𝑑𝑒𝑡[ 𝐴1 + 𝐴𝑛 , … , 𝐴𝑛−1 + 𝐴𝑛 , 𝐴𝑛 , 0𝑚 ]
= 𝑑𝑒𝑡[ 𝐴1 , … , 𝐴𝑛−1 , 0𝑚 , 0𝑚 − 𝐴𝑛 ]
= 𝑑𝑒𝑡[ 𝐴1 , … , 𝐴𝑛−1 , 𝐴𝑛 ]
55
1 2 7 4 3
Example 3.2.8 Let 𝐴 = [ ] , then det(𝐴 ) = 5
3 0 1 −1 2
det(𝐴2 ) = (−1)(3)(2) |2 7 4 3 1| = 5
0 1 −1 2 3
det(𝐴3 ) = (−1)(4)(2) |7 4 3 1 2| = 5
1 −1 2 3 0
1 2 7 2
Let 𝐵 = [ ], then det(𝐵 ) = 26
3 0 1 3
𝑚(𝑚+1)
Proof: Let 𝑟 = 1 + 2 + ⋯ + 𝑚 = and 𝐵𝑘 = 𝐴𝑛+1−𝑘 ,
2
(𝑚−1)𝑚
𝑘 ∈ {1,2, … , 𝑛}. Since exactly (𝑚 − 1) + (𝑚 − 2) + ⋯ + 1 =
2
inversions
of (adjacent) columns are needed to reverse the columns of a 𝑚 × 𝑚 matrix, we
have
|𝐵1 , 𝐵2 , … , 𝐵𝑛 | = ∑1≤𝑖1<⋯<𝑖𝑚 ≤𝑛(−1)𝑟+𝑖1+𝑖2+⋯+𝑖𝑚 |𝐵𝑖1 , 𝐵𝑖2 , … , 𝐵𝑖𝑚 |
56
(𝑚−1)𝑚
= ∑ (−1)𝑟+𝑖1 +𝑖2+⋯+𝑖𝑚+ 2 |𝐵𝑖𝑚 , 𝐵𝑖𝑚−1 , … , 𝐵𝑖1 |
1≤𝑖1 <⋯<𝑖𝑚 ≤𝑛
(𝑚−1)𝑚
= ∑ (−1)𝑟+𝑖1 +𝑖2+⋯+𝑖𝑚+ 2 |𝐴𝑛+1−𝑖𝑚 , 𝐴𝑛+1−𝑖𝑚−1 , … , 𝐴𝑛+1−𝑖1 |.
1≤𝑖1 <⋯<𝑖𝑚 ≤𝑛
(𝑚−1)𝑚
= ∑ (−1)𝑟+(𝑛+1)−(𝑗1 +𝑗2 +⋯+𝑗𝑚)+ 2 |𝐴𝑗1 , 𝐴𝑗2 , … , 𝐴𝑗𝑚 |
1≤𝑗1 <⋯<𝑗𝑚 ≤𝑛
(𝑚−1)𝑚
= ∑ (−1)𝑟+𝑚(𝑛+1)−(𝑗1 +𝑗2 +⋯+𝑗𝑚)+ 2 |𝐴𝑗1 , 𝐴𝑗2 , … , 𝐴𝑗𝑚 |
1≤𝑗1 <⋯<𝑗𝑚 ≤𝑛
(𝑚−1)𝑚
= |𝐴1 , 𝐴2 , … , 𝐴𝑛−1 , 𝐴𝑛 |. (−1)𝑚(𝑛+1)+ 2
𝑚(2𝑛+𝑚+1)
= |𝐴1 , 𝐴2 , … , 𝐴𝑛−1 , 𝐴𝑛 |. (−1) 2
3 4 7
|
2 1| = (−1)22(10+2+1) det(𝐴) = (−1)(5) = −5
2 −1 1 0 3
𝑚 ≡ 2(𝑚𝑜𝑑4) → 2 𝑚𝑜𝑑 4 = 2
3 4 7 2 1 1 2 7 4 3 (−1)
| |=| | = −5
2 −1 1 0 3 3 0 1 −1 2
57
Chapter Four
matrices.
First we state and prove the following result that relates the determinant of a
𝑢1 𝑣1
det [𝑢 𝑣2 ]
2
Proof: The parallelogram defined by the columns of the above matrix is the one
𝐵 = ‖𝑢
⃗ ‖, and 𝐻 2 + ‖𝑝𝑟𝑜𝑗𝑢 𝑣‖2 = ‖𝑣‖2
⃗ .𝑢
𝑣 ⃗ 2
𝐻2 = 𝑣. 𝑣 − ‖ ⃗ ⃗
𝑢‖ (see theorem 1.9.6)
𝑢.𝑢
58
𝑣. 𝑢
⃗ 𝑣. 𝑢
⃗
𝐻2 = 𝑣. 𝑣 − ( 𝑢
⃗ . 𝑢
⃗)
𝑢
⃗ .𝑢
⃗ 𝑢
⃗ .𝑢
⃗
(𝑣. 𝑢
⃗ )(𝑣. 𝑢
⃗)
𝐻2 = 𝑣. 𝑣 − ( (𝑢 ⃗ ))
⃗⃗⃗ . 𝑢
(𝑢 ⃗ )(𝑢
⃗ .𝑢 ⃗)
⃗ .𝑢
2
(𝑣. 𝑢⃗ )2
𝐻 = 𝑣. 𝑣 − ( )
(𝑢
⃗ .𝑢⃗)
Now, 𝐴2 = 𝐵 2 𝐻 2
2
⃗ )2
(𝑣. 𝑢
𝐴 =𝑢
⃗ .𝑢
⃗ (𝑣. 𝑣 − )
(𝑢 ⃗)
⃗ .𝑢
𝐴2 = (𝑢 ⃗ )2
⃗ )(𝑣. 𝑣) − (𝑣. 𝑢
⃗ .𝑢
So,
𝐴2 = 𝑢12 𝑣12 + 𝑢12 𝑣22 + 𝑢22 𝑣12 + 𝑢22 𝑣22 − 𝑢12 𝑣12 − 2𝑢1 𝑣1 𝑢2 𝑣2 − 𝑢22 𝑣22
𝑢1 𝑣1 2
𝐴2 = (𝑢1 𝑣2 − 𝑢2 𝑣1 )2 = (det [𝑢 𝑣2 ])
2
𝑢1 𝑣1
Area = |det [𝑢 𝑣2 ]|
2
The area of the triangle whose heads are 𝑢 = (𝑢1 , 𝑢2 ) 𝑎𝑛𝑑 𝑣 = (𝑣1 , 𝑣2 ) is
1 𝑢1 𝑣1
Area = 2 |det [𝑢 𝑣2 ]|
2
59
Also, the area of the quadrilateral can be found by dividing the shape into four
1 1 1 1
area 𝐴1 𝐴2 𝐴3 𝐴4 = |𝐴1 , 𝐴2 | + |𝐴2 , 𝐴3 | + |𝐴3 , 𝐴4 | + |𝐴4 , 𝐴1 |
2 2 2 2
This division of quadrilateral into triangles gives us the idea of calculating the
area of any polygon in 𝑅 2 .
1 𝑥1 𝑥2 𝑥2 𝑥3 𝑥𝑛 𝑥1
𝐴= (| | + | | + ⋯ + | 𝑦1 |)
2 𝑦1 𝑦2 𝑦2 𝑦3 𝑦𝑛
60
1
area of 𝐴1 … 𝐴𝑛 = 2 (|𝐴1 , 𝐴2 | + |𝐴2 , 𝐴3 | + ⋯ + |𝐴𝑛−1 , 𝐴𝑛 | + |𝐴𝑛 , 𝐴1 |).
Now, if 𝑛 = 3. The area 𝐴 with vertices (𝑥1 , 𝑦1 ), (𝑥2 , 𝑦2 ), (𝑥3 , 𝑦3 ) listed counter-
clockwise around the perimeter is given by
1 𝑥1 𝑥2 𝑥2 𝑥3 𝑥3 𝑥1
𝐴= (| 𝑦2 | + |𝑦2 𝑦3 | + |𝑦3 𝑦1 |)
2 𝑦1
1
= ((𝑥1 𝑦2 − 𝑥2 𝑦1 ) + (𝑥2 𝑦3 − 𝑥3 𝑦2 ) + (𝑥3 𝑦1 − 𝑥1 𝑦3 ))
2
1
= 2 (𝑥1 (𝑦2 − 𝑦3 ) − 𝑥2 (𝑦1 − 𝑦3 ) + 𝑥3 (𝑦1 − 𝑦2 ))
1 𝑥1 𝑥2 𝑥3
= 2 |𝑦 𝑦2 𝑦3 |
1
1
𝐴 = (|𝐴1 , 𝐴2 | + |𝐴2 , 𝐴3 | + ⋯ + |𝐴𝑛−1 , 𝐴𝑛 | + |𝐴𝑛 , 𝐴1 |)
2
= |𝐴1 , 𝐴2 , … , 𝐴𝑛 |
Proof: What we need here is to prove the second equality, we shall proceed by
1
showing that the formula (|𝐴1 , 𝐴2 | + |𝐴2 , 𝐴3 | + ⋯ + |𝐴𝑛−1 , 𝐴𝑛 | + |𝐴𝑛 , 𝐴1 |) is
2
1 1 0 0 1
A1: The area of a triangle with vertices (1,0), (0,1), (0,0) is | |=2
2 0 1 0
A2: It is clear that multiplying a single row by a constant 𝑐 means the dialation or
contraction of one coordinate while keeping the other coordinate fixed and in this
case new area equals |𝑐| times old area.
A3: A Let 𝐴 a polygon with vertices (𝑥1 , 𝑦1 ), (𝑥2 , 𝑦2 ), … , (𝑥𝑛 , 𝑦𝑛 ) such that
𝑥𝑖 = 𝑢𝑖 + 𝑣𝑖 , and 𝐵 a polygon with vertices (𝑢1 , 𝑦1 ), (𝑢2 , 𝑦2 ), … , (𝑢𝑛 , 𝑦𝑛 ), and
𝐶 a polygon with vertices (𝑣1 , 𝑦1 ), (𝑣2 , 𝑦2 ), … , (𝑣𝑛 , 𝑦𝑛 )
𝑢 + 𝑣 𝑢2 + 𝑣2 ⋯ 𝑢𝑛 + 𝑣𝑛
Suppose that, area of a polygon 𝐴 = [𝐴1 , … , 𝐴𝑛 ] = [ 1 𝑦 1 𝑦2 ⋯ 𝑦𝑛 ],
1
61
𝑢1 𝑢2 ⋯ 𝑢𝑛
but, area of a polygon 𝐵 = [𝐵1 , … , 𝐵𝑛 ] = [ 𝑦 𝑦2 ⋯ 𝑦𝑛 ],
1
𝑣1 𝑣2 ⋯ 𝑣𝑛
area of a polygon 𝐶 = [𝐶1 , … , 𝐶𝑛 ] = [ 𝑦 𝑦2 ⋯ 𝑦𝑛 ],
1
1
that is, area of 𝐴1 … 𝐴𝑛 = 2 (|𝐴1 , 𝐴2 | + |𝐴2 , 𝐴3 | + ⋯ + |𝐴𝑛−1 , 𝐴𝑛 | + |𝐴𝑛 , 𝐴1 |)
1 𝑥1 𝑥2 𝑥2 𝑥3 𝑥𝑛 𝑥1
= (| 𝑦2 | + |𝑦2 𝑦3 | + ⋯ + |𝑦𝑛 𝑦1 |)
2 𝑦1
1 𝑢1 + 𝑣1 𝑢2 + 𝑣2 𝑢2 + 𝑣2 𝑢3 + 𝑣3 𝑢𝑛 + 𝑣𝑛 𝑢1 + 𝑣1
= (| 𝑦 𝑦2 | + | 𝑦2 𝑦3 | + ⋯ + | 𝑦𝑛 𝑦1 |)
2 1
1 𝑢1 𝑢2 𝑣1 𝑣2 𝑢2 𝑢3 𝑣2 𝑣3 𝑢𝑛 𝑢1 𝑣𝑛 𝑣1
= ( |𝑦 𝑦2 | + | 𝑦1 𝑦2 | + | 𝑦2 𝑦3 | + | 𝑦2 𝑦3 | + ⋯ + | 𝑦𝑛 𝑦1 | + | 𝑦𝑛 𝑦1 |)
2 1
1 𝑢1 𝑢2 𝑢2 𝑢3 𝑢𝑛 𝑢1 1 𝑣1 𝑣2 𝑣2 𝑣3
= ( | 𝑦 𝑦2 | + | 𝑦2 𝑦3 | + ⋯ + | 𝑦𝑛 𝑦1 |) + (| | + | 𝑦3 | +
2 1 2 𝑦1 𝑦2 𝑦2
𝑣𝑛 𝑣1
|𝑦 𝑦1 |)
𝑛
area of 𝐴1 … 𝐴𝑛
1 1
= (|𝐵1 , 𝐵2 | + |𝐵2 , 𝐵3 | + ⋯ + |𝐵𝑛 , 𝐵1 |) + (|𝐶1 , 𝐶2 | + |𝐶2 , 𝐶3 | + ⋯ + |𝐶𝑛 , 𝐶1 |)
2 2
A4: Exchanging two rows geometrically means the reflection of the heads of the
polygon about in the straight line 𝑦 = 𝑥, That is, the image of each point (𝑥, 𝑦)
0 1 𝑥 𝑦
under the transformation is [ ] [𝑦] = [ ], which keeps the area fixed.
1 0 𝑥
62
Example 4.1.4 Use matrices to find the area of triangle with vertices:
1 4 7 2 13
area of triangle = | |= 2
2 0 2 3
1
area of 𝐴1 … 𝐴𝑛 = |𝐴1 + 𝐴2 , 𝐴2 + 𝐴3 , … , 𝐴𝑛−1 + 𝐴𝑛 , 𝐴𝑛 + 𝐴1 |
2
|𝐴1 + 𝐴2 , 𝐴2 + 𝐴3 , … , 𝐴𝑛−1 + 𝐴𝑛 , 𝐴𝑛 + 𝐴1 | =
From the adjacent figure it appears that the triangular shape resulted from the vertices shift
1
area of 𝐴1 𝐴2 𝐴3 = |𝐴1 + 𝐴2 , 𝐴2 + 𝐴3 , 𝐴3 + 𝐴1 |
2
1
= 2 (|𝐴1 + 𝐴2 , 𝐴2 + 𝐴3 | + |𝐴2 + 𝐴3 , 𝐴3 + 𝐴1 | + |𝐴3 + 𝐴1 , 𝐴1 + 𝐴2 |)
63
Applying Theorem 3.1.10 with 𝑋1 = −𝐴2 , 𝑋2 = −𝐴3 , 𝑋3 = −𝐴1 on the
determinants in order
1
area of 𝐴1 𝐴2 𝐴3 = 2 (|𝐴1 , 𝐴3 | + |𝐴2 , 𝐴1 | + |𝐴3 , 𝐴2 |)
1
= (|𝐴1 , 𝐴2 | + |𝐴2 , 𝐴3 | + |𝐴3 , 𝐴1 |)
2
|𝐴1 + 𝐴2 , 𝐴2 + 𝐴3 , … , 𝐴𝐾−1 + 𝐴𝐾 , 𝐴𝑘 + 𝐴1 |
= det[𝐴1 + 𝐴2 , 𝐴2 + 𝐴3 , … , 𝐴𝑘−1 + 𝐴𝑘 ]
+ det[𝐴𝑘+1 , 𝐴1 ]
64
Example 4.2.2 Use Theorem 4.2.1 to compute the area of the polygon in 𝑅 2 with
vertices (0, 0), (1, 0), (1, 1), (0 ,1)
1
Now, Area 𝐴1 𝐴2 𝐴3 𝐴4 = 2 |𝐴1 + 𝐴2 , 𝐴2 + 𝐴3 , 𝐴3 + 𝐴4 , 𝐴4 + 𝐴1 |
1 1 2 1 0
= | |=1
2 0 1 2 1
|𝐴1 + 𝑋, … , 𝐴𝑛 + 𝑋| = |𝐴1 , … , 𝐴𝑛 |
that is, the area of the polygon 𝐴1 𝐴2 … 𝐴𝑛 will not be changed when a fixed
vector 𝑋 = (𝑥, 𝑦) is subtracted from all the vertices (heads) of the polygon.
1 2 7 4 3 2 3 8 5 4
| |=| | = −3
3 0 1 −1 2 6 3 4 2 5
area of a polygon = 3
65
Theorem 4.2.5 [13]. Let 𝐴1 … 𝐴𝑛 be a polygon in 𝑅 2 with odd 𝑛.
Then
Example 4.2.6 Use Theorem 4.2.5 to find the area of a polygon with vertices:
3 5 7 7 3 5
| |=| |=8
2 1 4 4 2 1
|𝐴1 + 𝑋, … , 𝐴𝑛 + 𝑋| = |𝐴1 , … , 𝐴𝑛 |
Only if ∑𝑛𝑖=1(−1)𝑖 𝐴𝑖 = 0.
Proof: This theorem is incompatible with the theorem 3.1.10 and the reason is
𝑚 = 2, if 𝑛 is even, then 𝑚+𝑛 is even, This does not meet the theorem 3.1.10
requirement
66
Now, by Theorem 2.5.7, it is clear that for any point 𝑝 in 𝑅 2 holds
Only if 𝐴1 − 𝐴2 + ⋯ − 𝐴𝑛 = 0.
By Lemma 3.1.8
|𝐴1 , 𝐴2 , … , 𝐴𝑛 , 𝑝| = |𝐴1 + 𝑋, 𝐴2 + 𝑋, … , 𝐴𝑛 + 𝑋|
We get,
|𝐴1 , 𝐴2 , … , 𝐴𝑛 | = |𝐴1 + 𝑋, 𝐴2 + 𝑋, … , 𝐴𝑛 + 𝑋|
That is, the area of the polygon 𝐴1 𝐴2 … 𝐴𝑛 will not be changed when a fixed
vector 𝑋 = (𝑥, 𝑦) is subtracted from all the vertices (heads) of the polygon.
|𝐴1 , 𝐴2 , … , 𝐴𝑛 | = |0, 𝐴2 − 𝐴1 , … , 𝐴𝑛 − 𝐴1 |
= |𝐴1 − 𝐴2 , … , 𝐴1 − 𝐴𝑛 |
= |𝐵1 , 𝐵2 , … , 𝐵𝑛−1 |
1
= |𝐵 , 𝐵 | + |𝐵2 , 𝐵3 | + ⋯ + |𝐵𝑛−1 , 𝐵1 |
2 1 2
67
Example 4.2.8 Use Theorem 4.2.7 to find the area of a polygon with vertices:
Since −𝐴1 + 𝐴2 − 𝐴3 + 𝐴4 = 0
area of a polygon
1 3 4 2
𝐴1 𝐴2 𝐴3 𝐴4 = || || = |−9 | = 9
2 5 2 −1
1 2 4 3
area of a polygon 𝐴1 𝐴4 𝐴3 𝐴2 = | |=9
2 −1 2 5
2 4 5 3
= || || = |−9 | = 9
4 7 4 1
= |02 , 𝐴𝑛 − 𝐴1 , … , 𝐴𝑛 − 𝐴𝑛−1 |
68
Adding −𝐴𝑛 to each column
Take out a common negative signal factor from the two rows
Note that, this says that we can calculate the area by starting from any vertex of
the polygon paying attention to the direction and arrangement of the vertices.
Example 4.2.10 Use Theorem 4.2.9 to find the area of a polygon with vertices:
Since −𝐴1 + 𝐴2 − 𝐴3 + 𝐴4 = 0
1 3 4 2 2 1 3 4
|| || = || || = 9
2 5 2 −1 −1 2 5 2
69
We notice that if we displace the shape to the original point and calculate the
area by dividing the shape into such based on the origin point, the area does not
change.
𝑑𝑒𝑡[ 𝐴1 , 𝐴2 , 𝐴3 , 𝐴4 ] = 𝑑𝑒𝑡[ 𝐴1 , 𝐴2 , 𝐴3 ]
Now, | 𝐴1 , 𝐴2 , 𝐴3 | = | 𝐴1 , 𝐴2 | − | 𝐴1 , 𝐴3 | + | 𝐴2 , 𝐴3 |
= | 𝐴1 , 𝐴2 | + |−𝐴1 + 𝐴2 , 𝐴3 |
| 𝐴1 , 𝐴2 , 𝐴3 | = | 𝐴1 , 𝐴2 | + |𝐴3 − 𝐴4 , 𝐴3 |
= | 𝐴1 , 𝐴2 | + |− 𝐴4 , 𝐴3 | = | 𝐴1 , 𝐴2 | + | 𝐴3 , 𝐴4 |
| 𝐴1 , 𝐴2 , 𝐴3 , 𝐴4 | = | 𝐴1 , 𝐴2 | + | 𝐴3 , 𝐴4 |
Example 4.2.13 Use Theorem 4.2.11 to find the area of a polygon with vertices:
Since −𝐴1 + 𝐴2 − 𝐴3 + 𝐴4 = 0
|𝐴1 , 𝐴2 , 𝐴3 , 𝐴4 | = |𝐴1 − 𝐴4 , 𝐴2 − 𝐴4 , 𝐴3 − 𝐴4 , 𝐴4 − 𝐴4 |
= |𝐴1 − 𝐴4 , 𝐴2 − 𝐴4 , 𝐴3 − 𝐴4 , 0| = |𝐴1 − 𝐴4 , 𝐴2 − 𝐴4 , 𝐴3 − 𝐴4 |
−1 1 2 0 −1 1 2
=| |=| | = −9
3 6 3 0 3 6 3
area of a polygon 𝐴1 𝐴2 𝐴3 𝐴4 = 9
70
Note that the point can be considered as the center on which to divide triangles to
calculate the area
then,
1 3 4 2 1 3 4
area of a polygon 𝐴1 𝐴2 𝐴3 𝐴4 = || || = || || = |−9| = 9
2 5 2 −1 2 5 2
∑𝑛𝑖=1(−1)𝑖 𝐴𝑖 = 0. Then
Proof:
area of 𝐴1 … 𝐴𝑛
1
= 2 (|𝐴1 , 𝐴2 | + |𝐴2 , 𝐴3 | + ⋯ + |𝐴𝑘−1 , 𝐴𝑘 | + |𝐴𝑘 , 𝐴𝑘+1 | + |𝐴𝑘+1 , 𝐴𝑘+2 | + ⋯ +
|𝐴𝑛−1 , 𝐴𝑛 | + |𝐴𝑛 , 𝐴1 |)
1
= 2 (|𝐴1 , 𝐴2 | + |𝐴2 , 𝐴3 | + ⋯ + |𝐴𝑘−1 , 𝐴𝑘 | + |𝐴𝑘+1 , 𝐴𝑘+2 | + ⋯ + |𝐴𝑛−1 , 𝐴𝑛 |) +
1
(|𝐴𝑘 , 𝐴𝑘+1 | + |𝐴𝑛 , 𝐴1 |)
2
1
|𝐴1 , … , 𝐴𝑛 | = |𝐴1 , … , 𝐴𝑘 | + |𝐴𝑘+1 , … , 𝐴𝑛 | + (|𝐴𝑘 , 𝐴𝑘+1 | + |𝐴𝑛 , 𝐴1 |) −
2
1
(|𝐴𝑘 , 𝐴1 | + |𝐴𝑛 , 𝐴𝑘+1 |)
2
1
let ∆ = 2 (|𝐴𝑘 , 𝐴𝑘+1 | + |𝐴𝑛 , 𝐴1 | − |𝐴𝑘 , 𝐴1 | − |𝐴𝑛 , 𝐴𝑘+1 |)
71
Example 4.2.15 Use Theorem 4.2.14 to find the area of a polygon with vertices:
Since
−𝐴1 + 𝐴2 − 𝐴3 + 𝐴4 + 𝐴5 − 𝐴6 = 0
Then,
area of a polygon 𝐴1 𝐴2 𝐴3 𝐴4 𝐴5 𝐴6 =
1 0 1 2 3 3
|| || = 4
2 1 0 −1 −1 1
take 𝑘 = 3 then
take 𝑘 = 2 then
Then
72
𝐴1 +𝐴2 𝐴2 +𝐴3 𝐴𝑛 +𝐴1
Proof: 4|𝐵1 , … , 𝐵𝑛 | = 4 | , ,…, |.
2 2 2
By theorem 4.2.1
1 1 1
4|𝐵1 , … , 𝐵𝑛 | = 4 ( |𝐴1 , 𝐴2 | + |𝐴2 , 𝐴3 | + ⋯ + |𝐴𝑛 , 𝐴1 |)
4 4 4
1
area of a polygon 𝐵1 𝐵2 𝐵3 𝐵4 = area of a polygon 𝐴1 𝐴2 𝐴3 𝐴4
2
Since −𝐴1 + 𝐴2 − 𝐴3 + 𝐴4 = 0
Since −𝐵1 + 𝐵2 − 𝐵3 + 𝐵4 = 0
2 3.5 3 1.5
Then, area of a polygon 𝐵1 𝐵2 𝐵3 𝐵4 = || || = 4.5
3.5 3.5 0.5 0.5
1
area of a polygon 𝐵1 𝐵2 𝐵3 𝐵4 = area of a polygon 𝐴1 𝐴2 𝐴3 𝐴4 .
2
73
Chapter Five
In linear algebra, the inverse of a matrix is defined only for square matrices,
and if a matrix is singular, it does not have an inverse.
The aim of this chapter is the discussion of existence of inverses for non-
square matrices. We know that the fundamental idea for existence of inverse of
matrix it must be nonsingular. (it has non-zero determinant).
(i) A non-singular non-square matrix 𝐴 has a left inverse if there exists a matrix
𝐴−1
𝐿 such that 𝐴−1
𝐿 𝐴 = 𝐼, where 𝐼 denote the identity matrix.
(ii) A non-singular non-square matrix 𝐴 has a right inverse if there exists a matrix
𝐴−1
𝑅 such that 𝐴 𝐴−1
𝑅 = 𝐼, where 𝐼 denote the identity matrix.
1 3 0
Example 5.1.2 Let A = [ ] Find a right inverse of A
2 2 1
𝑥1 𝑦1
𝐴 𝐴−1
𝑅 = [
1 3 0 𝑥
] [ 2 𝑦2 ] = [1 0]
2 2 1 𝑥 𝑦3 0 1
3
𝑥1 + 3𝑥2 𝑦1 + 3𝑦2 1 0
[ ]=[ ]
2𝑥1 + 2𝑥2 + 𝑥3 2𝑦1 + 2𝑦2 + 𝑦3 0 1
74
Gauss-Jordan reduction Using we obtain the following
𝑥1 + 34 𝑥3 = −1
2
, 𝑥2 − 14 𝑥3 = 12 , 𝑦1 + 34 𝑦3 = 34 , 𝑦2 − 14 𝑦3 = −1
4
(*)
This system has infinitely many solutions, one solution gives the following
1 1 0 0
−1⁄ 13
𝐴−1
𝑅 = [0 3] , another solution is 𝐴−1
𝑅 = [
⁄32 0]
−2 −1⁄ −3⁄ 1
3 8
Note: When a right inverse or a left inverse for a non-singular non-square matrix
exists, it is not unique.
Remark 5.1.3 For any non-singular square matrix 𝐴, left inverse and right inverse
exists and it is equal to inverse of 𝐴, that is
1
𝐴−1 −1
𝑅 = 𝐴𝐿 = 𝐴
−1
= det(𝐴) 𝑎𝑑𝑗(𝐴).
The following theorem is from [3], but we give here another proof
Theorem 5.1.4 Every non-singular horizontal matrix 𝐴 has a right inverse 𝐴−1
𝑅 given
by
1
𝐴−1
𝑅 = 𝑎𝑑𝑗(𝐴)
det(𝐴)
Proof:
𝑎11 𝑎12 ⋯ 𝑎1𝑛
𝑎21 𝑎22 ⋯ 𝑎2𝑛 𝑀11 𝑀21 ⋯ 𝑀𝑗1 ⋯ 𝑀𝑚1
⋮ ⋮ ⋱ ⋮ 𝑀 𝑀22 … 𝑀𝑗2 ⋯ 𝑀𝑚2
Let 𝐴 = 𝑎𝑖1 𝑎𝑖2 ⋯ 𝑎𝑖𝑛 , 𝑎𝑑𝑗(𝐴) = [ 12 ]
⋮ ⋮ ⋱ ⋮ ⋱ ⋮
⋮ ⋮ ⋱ ⋮ 𝑀1𝑛 𝑀2𝑛 … 𝑀𝑗𝑛 ⋯ 𝑀𝑚𝑛
[𝑎𝑚1 𝑎𝑚2 ⋯ 𝑎𝑚𝑛 ]
Since 𝐴 is non-singular, 𝑑𝑒𝑡(𝐴) is non-zero the matrix
75
𝑀11 𝑀21 ⋯ 𝑀𝑗1 ⋯ 𝑀𝑚1
1 1 𝑀 𝑀22 … 𝑀𝑗2 ⋯ 𝑀𝑚2
𝐵= 𝑎𝑑𝑗(𝐴) = [ 12 ]
det(𝐴) det(𝐴) ⋮ ⋮ ⋱ ⋮ ⋱ ⋮
𝑀1𝑛 𝑀2𝑛 … 𝑀𝑗𝑛 ⋯ 𝑀𝑚𝑛
satisfies
and, 𝑖𝑓 𝑖 ≠ 𝑗, we need to show that 𝑎𝑖1 𝑀𝑗1 + 𝑎𝑖2 𝑀𝑗2 + ⋯ + 𝑎𝑖𝑛 𝑀𝑗𝑛 = 0
det(𝐴) 0 ⋯ 0
1 0
𝐴𝐵 = [ 0 det(𝐴) ⋯
⋮ ] = 𝐼𝑚
det(𝐴) ⋮ ⋮ ⋱
0 0 ⋯ det(𝐴)
Hence, the matrix 𝐵 is a right inverse of 𝐴.
1 3 0
In example 5.1.2 we have seen that a right inverse for A= [ ] is
2 2 1
1 1
−1
𝐴−1
𝑅 = [0 ⁄3] here we compute an inverse of the same 𝐴 by theorem 5.1.4
−2 −1⁄3
−1⁄ 3⁄
2 2
which is 𝐴−1
𝑅 = [ 1⁄ −1⁄ ] . it is easy to see that the resulting inverse is
2 2
0 1
another solution to the system (*).
76
1 1 2 0
Example 5.1.5 Let 𝐴 = [ 1 2 1 2]
3 4 1 2
−1⁄ −5⁄ 1⁄
2 4 4
2 5 −1 1 3⁄ 1⁄
1 1 −2 −3 −1 ⁄2 4 4
𝐴−1
𝑅 = 𝑎𝑑𝑗(𝐴) = [ ]=
det(𝐴) −4 −2 −1 1 1⁄ 1⁄ −1⁄
2 −1 1 2 4 4
−1⁄ 1⁄ −1⁄
[ 2 4 4]
Theorem 5.1.6 [3 ] Every non-singular vertical matrix 𝐴 has a left inverse 𝐴−1
𝐿 , such
that
1
𝐴−1
𝐿 = 𝑎𝑑𝑗(𝐴)
det(𝐴)
The following theorem describes the case that 𝐴 has both a right inverse and a left inverse.
If 𝐴 𝐴−1
𝑅 = 𝐼𝑚 , then the equation 𝑨𝒙 = 𝒃 has a solution for every possible 𝒃 in 𝑅
𝑚
Let 𝐴−1
𝑅 = [𝑟1 ⋯ 𝑟𝑚 ]. Then
𝐴𝐴−1
𝑅 = 𝐴 ⋅ [𝑟1 ⋯ 𝑟𝑚 ] = [𝐴 𝑟1 ⋯ 𝐴 𝑟𝑚 ] = [𝑒1 ⋯ 𝑒𝑚 ] = 𝐼𝑚 .
77
Therefore 𝐴 has a pivot position in every row. This force 𝑚 ≤ 𝑛, since every
pivot position must be in a different column.
If 𝐴−1 −1 −1
𝐿 𝐴 = 𝐼𝑛 , consider the equation 𝑨𝒙 = 0. Then 𝐴𝐿 𝑨 𝒙 = 𝐴𝐿 0 = 0.
But 𝐴−1
𝐿 𝑨 𝒙 = 𝐼𝑛 𝒙 = 𝒙, so 𝒙 = 𝟎. In other words, 𝑨𝒙 = 𝟎 has a unique solution
and therefore the columns of 𝐴 must be linearly independent (see definition 1.9.3)
and therefore each column must be a pivot position column. Since each pivot
position must be in a different row, these forces 𝑛 ≤ 𝑚.
So, combining the two paragraphs gives that 𝑚 = 𝑛. Since 𝐴 is now known to be
square, 𝐴 is invertible and 𝐴 −1 = 𝐴−1 −1
𝑅 = 𝐴𝐿 .
In Chapter one we have known that taking inverse and transposing a matrix
commute. (theorem 1.5.2). Here we write an example that asserts this fact for a
non-square matrix.
−1⁄ 3⁄
1 3 0 2 2
Example 5.2.1 Let A = [ ], a right inverse of A is A−1
𝑅 = [ 1⁄ −1⁄ ],
2 2 1 2 2
0 −1
−1⁄ 1⁄ 0 1 2
(A−1 𝑇 2 2 AT = [3 2] ,
𝑅 ) = [
3⁄ −1⁄ −1
], also
2 2 0 1
−1⁄ 1⁄ 0
(AT )−1 2 2 ] = (A−1 𝑇
𝐿 =[
3⁄ −1⁄ 𝑅 )
2 2 −1
(AT )−1 −1 𝑇
𝐿 = (A𝑅 ) and (AT )−1 −1 𝑇
𝑅 = (A𝐿 )
A−1 −1
𝑅 is a right inverse of A, we have 𝐴 𝐴𝑅 = I𝑚
78
Taking transposes, we obtain
(𝐴 𝐴−1 𝑇
𝑅 ) = (I𝑚 )
𝑇
(A−1 𝑇 𝑇
𝑅 ) 𝐴 = I𝑚
(AT )−1 −1 𝑇
𝐿 = (A𝑅 )
(A B)−1 −1 −1
𝑅 = B𝑅 𝐴 and (A B)−1 −1 −1
𝐿 = B𝐿 𝐴
Now,
Therefore, (A B)−1 −1 −1
𝑅 = B𝑅 𝐴 and (A B)−1 −1 −1
𝐿 = B𝐿 𝐴
1 3 1 3 0
Example 5.2.4 Prove theorem 5.2.3 for the matrices 𝐴 = [ ] , 𝐵= [ ]
2 1 2 2 1
7 −6
7 9 3
Then, 𝐵 = [ ], 𝑎𝑑𝑗( 𝐴 𝐵 ) = [−3 4]
4 8 1
−4 2
7⁄ −6⁄
10 10 −1⁄ 3⁄
(A B)−1 = −3 ⁄10 4⁄ and A−1
= [ 5 5]
𝑅 10 2⁄ −1⁄
−4 2⁄ 5 5
[ ⁄10 10 ]
−1⁄ 3⁄
1 −3 2 2
𝑎𝑑𝑗(𝐵) = [−1 1 ], B𝑅−1 = [ 1⁄ −1⁄ ]
0 2 2 2
0 −1
79
7⁄ −6⁄
10 10
B𝑅−1 𝐴−1 −3 4⁄ −1
= ⁄10 10 = (A B)𝑅
−4 2⁄
[ ⁄10 10 ]
1 −1
(k A)−1 = A
𝑘
1
(k A) ( 𝐴−1 ) = A 𝐴−1
𝑅 = 𝐼𝑚 , which gives
𝑘 𝑅
1 −1
(k A)−1
𝑅 = 𝐴
𝑘 𝑅
1
Also, if 𝑛 < 𝑚, (𝑘 𝐴−1 −1
𝐿 ) (k A) = 𝐴𝐿 A = 𝐼𝑛 which gives
1
(k A)−1 −1
𝐿 = 𝑘 𝐴𝐿
We notice here that the properties of the inverse that are satisfied in square
matrices (see theorem 1.5.2.b-c-d) are also satisfied in non-square matrices (see
theorem 5.2.2, 5.2.3, 5.2.5)
The following theorem is given in [8], but we are providing our own proof.
A adj(A) = det(𝐴) 𝐼𝑚
80
𝑎11 𝑎12 ⋯ 𝑎1𝑛
𝑎21 𝑎22 ⋯ 𝑎2𝑛 𝑀11 𝑀21 ⋯ 𝑀𝑗1 ⋯ 𝑀𝑚1
⋮ ⋮ ⋱ ⋮ 𝑀12 𝑀22 … 𝑀𝑗2 ⋯ 𝑀𝑚2
𝐴 𝑎𝑑𝑗(𝐴) = 𝑎𝑖1 𝑎𝑖2 ⋯ 𝑎𝑖𝑛 ⋱
⋮ ⋮ ⋱ ⋮ ⋮
⋮ ⋮ ⋱ ⋮ [𝑀1𝑛 𝑀2𝑛 … 𝑀𝑗𝑛 ⋯ 𝑀𝑚𝑛 ]
[𝑎𝑚1 𝑎𝑚2 ⋯ 𝑎𝑚𝑛 ]
and, 𝑖𝑓 𝑖 ≠ 𝑗, 𝑎𝑖1 𝑀𝑗1 + 𝑎𝑖2 𝑀𝑗2 + ⋯ + 𝑎𝑖𝑛 𝑀𝑗𝑛 = 0 )It was pre-established in
theorem 5.1.4 **)
det(𝐴) 0 ⋯ 0
0
Hence, 𝐴 𝑎𝑑𝑗(𝐴) = [ 0 det(𝐴) ⋯
⋮ ] = det(𝐴) 𝐼𝑚
⋮ ⋮ ⋱
0 0 ⋯ det(𝐴)
We note that this property is valid for square matrices except that the switch
1 3 0
Example 5.2.7 Prove theorem 5.2.6 for the matrices 𝐴 = [ ]
2 1 −1
2 3
Then, 𝑎𝑑𝑗(𝐴) = [3 1 ] , and 𝑑𝑒𝑡(𝐴) = −7
1 −2
All of the following corollaries in this section are proved by using theorem 5.2.6
as follows:
So, A adj(A) = 0,
If 𝐴 = 0 (null and singular), then 𝑎𝑑𝑗(𝐴) = 0 and hence 𝑎𝑑𝑗(𝐴) is singular too.
81
If 𝐴 ≠ 0 (non-null and singular), then 𝐴 contains a non-null row, say the 𝑖 𝑡ℎ row
𝑎 ′𝑖 it follows that 𝑎 ′𝑖 𝑎𝑑𝑗(𝐴) = 0
Which implies that the rows of 𝑎𝑑𝑗(𝐴) are linearly dependent (see definition
1.9.3), and hence 𝑎𝑑𝑗(𝐴) is singular.
1 1 1
Example 5.2.9 Prove theorem 5.2.8 for the matrices 𝐴 = [ ]
2 3 4
−1 0
then 𝑎𝑑𝑗(𝐴) = [ 2 0] , and 𝑑𝑒𝑡(𝐴) = 0, det(𝑎𝑑𝑗(𝐴)) = 0
−1 0
(adj(A))T = adj(AT )
𝑇
( adj(A)) = (det(𝐴) 𝐴−1
𝑅 )
𝑇
(adj(A))T = adj(AT )
This property is valid for singular matrices of all size ( square or non-square)
1 3 0
Example 5.2.11 Let 𝐴 = [ ]
2 1 −1
2 3
(𝑎𝑑𝑗(𝐴) )𝑇 = [2 3 1
then 𝑎𝑑𝑗(𝐴) = [3 1 ] , ]
3 1 −2
1 −2
82
1 2
2 3 1
A𝑇 = [3 1 ], 𝑎𝑑𝑗(A𝑇 ) = [ ] = (𝑎𝑑𝑗(𝐴) )𝑇
3 1 −2
0 −1
1 −1
adj(k A) = 𝑑𝑒𝑡(k A)(𝑘 𝐴)−1 𝑚
𝑅 = 𝑘 det(𝐴) 𝐴
𝑘 𝑅
Similarly, if 𝑛<𝑚 .
This property is valid for singular matrices of all size ( square or non-square)
= adj(B). adj(A)
Similarly, if 𝑛<𝑚
1 3 1 3 0
Example 5.2.14 Let 𝐴 = [ ] , 𝐵= [ ]
2 1 2 2 1
83
7 −6
7 9 3
then 𝐴 𝐵 = [ ], 𝑎𝑑𝑗( 𝐴 𝐵 ) = [−3 4 ]
4 8 1
−4 2
1 −3
1 −3
𝑎𝑑𝑗(𝐵) = [−1 1 ], 𝑎𝑑𝑗(𝐴) = [ ]
−2 1
0 2
1 −3 7 −6
1 −3
𝑎𝑑𝑗(𝐵) 𝑎𝑑𝑗(𝐴) = [−1 1 ][ ] = [−3 4 ] = 𝑎𝑑𝑗( 𝐴 𝐵)
−2 1
0 2 −4 2
1 2 2 5 8
Example 5.2.15 Let 𝐴= [ ], 𝐵= [ ]
2 4 3 2 1
8 9 10
Then, det(𝐴) = 0, 𝑑𝑒𝑡(𝐵) = 0, 𝐴𝐵 =[ ], det(𝐴𝐵) = −4,
16 18 20
−2 1 1 3
) 4 −2
𝑎𝑑𝑗( 𝐴 𝐵 = [ 4 −2], 𝑎𝑑𝑗(𝐵) = [−2 −6], 𝑎𝑑𝑗(𝐴) = [ ]
−2 1
−2 −1 1 3
−2 1
𝑎𝑑𝑗(𝐵) 𝑎𝑑𝑗(𝐴) = [ 4 −2] = 𝑎𝑑𝑗( 𝐴 𝐵)
−2 −1
We notice here that the properties of the adjoint of a matrix that are satisfied
in square matrices (see theorem 1.4.4.d-e-f) are also satisfied in non-square
matrices (see theorem 5.2.10, 5.2.12, 5.2.13).
denoted by Ɲ(𝐴). The dimension of the null space of 𝐴 is called the nullity of 𝐴.
84
(1) Ɲ(𝐴) = Ɲ(𝐴𝑇 𝐴)
(2) 𝑟𝑎𝑛𝑘(𝐴) = 𝑟𝑎𝑛𝑘(𝐴𝑇 𝐴)
𝐴𝑇 𝑨𝒙 = 𝐴𝑇 0 = 0
Thus 𝑥 ∈ Ɲ(𝐴𝑇 𝐴)
𝑥 𝑇 𝐴𝑇 𝑨𝒙 = 𝑥 𝑇 0 = 0
and the length of the vector 𝑨𝒙 is zero, thus the vector 𝑨𝒙 = 𝟎. Hence 𝑥 ∈ Ɲ(𝐴)
(2) We use the rank-nullity theorem and obtain (see theorem 1.9.8)
𝑟𝑎𝑛𝑘(𝐴) = 𝑛 − 𝑑𝑖𝑚(Ɲ(𝐴))
= 𝑛 − 𝑑𝑖𝑚(Ɲ(𝐴𝑇 𝐴))
= 𝑟𝑎𝑛𝑘(𝐴𝑇 𝐴)
85
Definition 5.3.2 [4] The matrix 𝐴+ 𝑇 𝑇 −1
𝑅 = 𝐴 (𝐴 𝐴 ) , when 𝐴 is 𝑚 × 𝑛 (𝑚 ≤ 𝑛)
and (𝐴 𝐴𝑇 )−1 exists, and 𝑟𝑎𝑛𝑘 (𝐴) = 𝑚, is called the right Pseudo inverse of 𝐴.
and ( 𝐴𝑇 𝐴)−1 exists, and 𝑟𝑎𝑛𝑘 (𝐴) = 𝑛, is called the left Pseudo inverse of 𝐴
Remark 5.3.4 The matrix 𝐴𝑚×𝑛 ( 𝑚 > 𝑛) has 𝑟𝑎𝑛𝑘 = 𝑛, and therefore 𝐴𝑇 also
𝑛 × 𝑛 matrix. it therefore has full rank and its inverse exists. (see theorem 1.9.14)
1 1 1
Example 5.3.5 Let A = [ ], 𝑟𝑎𝑛𝑘 (𝐴) = 2 = 𝑚 = 𝑟𝑎𝑛𝑘 (𝐴𝐴𝑇 )
3 −1 1
11 −3
𝐴𝐴𝑇 is invertible, |𝐴 𝐴𝑇 | = 33 − 9 = 24, and (𝐴 𝐴𝑇 )−1 = 24
1
[ ]
−3 3
1 1 3
11 −3 1 2 6
𝐴−1 + 𝑇 𝑇 −1
𝑅 = 𝐴𝑅 = 𝐴 (𝐴 𝐴 ) = [1 −1] [ ]= [14 −6]
24 −3 3 24
1 1 8 0
1 1 1 1 2 6
𝐴 𝐴−1
𝑅 = [ ] [14 −6] = 𝐼2
24 3 −1 1
8 0
1 −2
Example 5.3.6 Let A = [−2 1 ], 𝑟𝑎𝑛𝑘 (𝐴) = 2 = 𝑛 = 𝑟𝑎𝑛𝑘 (𝐴𝑇 𝐴)
1 1
6 3
𝐴𝑇 𝐴 is invertible, | 𝐴𝑇 𝐴 | = 36 − 9 = 27, and ( 𝐴𝑇 𝐴)−1 = 27
1
[ ]
3 6
−1 1
1 6 0
3 1 −2 1 3 3]
𝐴−1
𝐿 = 𝐴+ 𝑇 −1 𝑇
𝐿 = ( 𝐴 𝐴) 𝐴 = [ ][ ][
27 3 6 −2 1 1 −1 1
0
3 3
−1 1
0 1 −2
3 3
𝐴−1
𝐿 𝐴 = [−1 1] [−2 1 ] = 𝐼2
3
0 3 1 1
86
Chapter Six
The solution of the system can be expressed as 𝒙 = 𝐴−1 𝒃 where 𝐴−1 is the
solution to the system, we will now try to find solution to the system.
pseudo-inverse of 𝐴
𝒙 = 𝐴−1 𝑇 −1 𝑇
𝐿 (𝐴 ) 𝐴 𝒃
𝒙 = 𝐴−1
𝐿 𝒃
87
Example 6.1.2 Find a solution to the system
𝑥1 + 3𝑥2 = −2
3 𝑥1 − 𝑥2 = 4
2 𝑥1 + 2 𝑥2 = 0
1 3 −2
𝑨=[3 −1] is the coefficient matrix, and 𝒃 = [ 4 ] is the constant vector
2 2 0
𝒙 = ( 𝐴𝑇 𝐴)−1 𝐴𝑇 𝐛
1 3
1 3 2 14 4
𝐴𝑇 𝐴 = [ ][ 3 −1] = [ ]
3 −1 2 4 14
2 2
1 7 −2
So, (𝐴𝑇 𝐴)−1 = 90 [ ]
−2 7
1 7 −2 1 3 2 1 2 23 10
( 𝐴𝑇 𝐴)−1 𝐴𝑇 = [ ][ ]= [ ]
90 −2 7 3 −1 2 90 19 −13 10
1 2 −2 1 90
𝑇 −1 𝑇 23 10 1
𝒙 = ( 𝐴 𝐴) 𝐴 𝐛 = [ ][ 4 ] = [ ]= [ ]
90 19 −13 10 90 −90 −1
0
Note that x = ( 1 , −1) represents the intersection point of the three straight lines
shown in the graph.
Note that : if 𝐴𝑇 𝐴 is invertble, then the only possibility for 𝑨𝒙 = 𝒃 are either unique
solution or no solution.
88
Example 6.1.3 Find a solution to the system
𝑥1 + 3 𝑥2 = 5
𝑥1 − 𝑥2 = 1
𝑥1 + 𝑥2 = 0
1 3 5
𝐴 = [1 −1] is the coefficient matrix and 𝒃 = [1] is the constant vector
1 1 0
𝒙 = ( 𝐴𝑇 𝐴)−1 𝐴𝑇 𝐛
1 3
1 1 1 3 3
We calculate 𝐴𝑇 𝐴 = [ ][ 1 −1] = [ ]
3 −1 1 3 11
1 1
1 11 −3
Next, (𝐴𝑇 𝐴)−1 = 24 [ ]
−3 3
1 11 −3 1 1 1 1 2 14 8
( 𝐴𝑇 𝐴)−1 𝐴𝑇 = [ ][ ]= [ ]
24 −3 3 3 −1 1 24 6 −6 0
5
𝑇 −1 𝑇 1 2 14 8 1 24 1
𝒙 = ( 𝐴 𝐴) 𝐴 𝐛 = [ ] [1 ] = [ ]= [ ]
24 6 −6 0 24 24 1
0
aright pseudo-inverse of 𝐴
89
and we can multiply 𝑨 𝒙 = 𝒃 by (𝐴 𝐴𝑇 )−1 on both sides, obtaining
(𝐴 𝐴𝑇 )−1 𝑨 𝒙 = (𝐴 𝐴𝑇 )−1 𝒃
(𝐴𝑇 )−1 −1 𝑇 −1
𝑅 𝐴𝐿 𝑨 𝒙 = (𝐴 𝐴 ) 𝒃
𝑇 −1
(𝐴 )𝑅 𝐼𝑛 𝒙 = 𝑇 −1
(𝐴 𝐴 ) 𝒃 (1)
𝒙 = 𝐴𝑇 (𝐴 𝐴𝑇 )−1 𝐛
2𝑥1 + 3 𝑥2 − 2𝑥3 = 4
−6 𝑥1 − 8 𝑥2 + 6𝑥3 = 1
𝒙 = ( 𝐴𝑇 𝐴)−1 𝐴𝑇 𝐛
2 −6
2 3 −2 17 −48
We calculate 𝐴 𝐴𝑇 = [ ] [ 3 −8] = [ ]
−6 −8 6 −48 136
−2 6
1 136 48
Next, (𝐴𝑇 𝐴)−1 = 8 [ ]
48 17
1 2 −6 1 −16 −6
136 48
𝐴𝑇 (𝐴 𝐴𝑇 )−1 = [ 3 −8] [ ] = [ 24 8]
8 48 17 8
−2 6 16 6
−70
−16 −6 −70 8
1 4 1
𝒙 = 𝐴𝑇 (𝐴 𝐴𝑇 )−1 𝐛 = 8 [ 24 8 ][ ] = [ 104 ] = [ 13 ] is a solution
1 8 70
16 6 70
8
−35
𝑥1 4 1
The general solution in vector form as [ 𝑥2 ] = [ 13 ] + 𝑡 [0]
𝑥3 35 1
4
This is a parameter vector equation of the line of intersection 𝐿 of the three lines. The
coordinates of each of 𝐿′𝑠 points make one of the infinitely many solutions of the system.
90
Theorem 6.1.6 [6 ]. (General solution)
𝒙 = 𝐴−1 −1
𝑅 𝒃 + (𝐼𝑛 − 𝐴𝑅 𝐴)𝑦, (1)
𝑨𝒙 = 𝐴 𝐴−1 −1
𝑅 𝒃 + 𝐴 (𝐼 − 𝐴𝑅 𝐴)𝑦
= 𝐼𝑚 𝒃 + (𝐴 − 𝐴 𝐴−1
𝑅 𝐴)𝑦 by hypothesis
= 𝒃 , since 𝐴 𝐴−1
𝑅 𝐴 = 𝐼𝑚 𝐴𝑚×𝑛 = 𝐴
𝑧 = 𝐴−1 −1 −1 −1
𝑅 𝑨 𝒛 + (𝐼 − 𝐴𝑅 𝐴)𝑧 = 𝐴𝑅 𝑨 𝒛 + 𝑧 − 𝐴𝑅 𝑨 𝒛
Note that: when 𝐴 is not full rank, then the above theorem cannot be used.
91
Let 𝐴𝑚×𝑛 , 𝑚 ≤ 𝑛, be nonsingular, then 𝐴−1
𝑅 exists and we can multiply 𝑨 𝒙 = 𝒃
by 𝐴−1 −1 −1
𝑅 on both sides, obtaining 𝑨 𝐴𝑅 𝒙 = 𝐴𝑅 𝒃
1
Then 𝐼𝑚 𝒙 = 𝐴−1
𝑅 𝒃 , but 𝐴−1
𝑅 = det(𝐴) 𝑎𝑑𝑗(𝐴) ( see theorem 5.1.4)
1
So, 𝒙 = 𝑎𝑑𝑗(𝐴) 𝒃
det(𝐴)
sides of 𝑨 𝒙 = 𝒃 by 𝐴−1
𝐿 , obtaining 𝐴−1 −1
𝐿 𝑨 𝒙 = 𝐴𝐿 𝒃
1
That is 𝐼𝑛 𝒙 = 𝐴−1
𝐿 𝒃, but 𝐴−1
𝐿 = det(𝐴) 𝑎𝑑𝑗(𝐴) (see theorem 5.1.6)
1
So, 𝒙 = 𝑎𝑑𝑗(𝐴) 𝒃
det(𝐴)
1 3 −2
𝐴 = [3 −1] is the coefficient matrix, and 𝒃 = [ 4 ] is the constant vector
2 2 0
1
𝑑𝑒𝑡(𝐴) = 2, since 𝐴 is 3 × 2, 𝐴−1
𝐿 exists, 𝒙 = 𝑎𝑑𝑗(𝐴) 𝒃
det(𝐴)
−3 −1 4
So, we calculate 𝑎𝑑𝑗(𝐴) = [ ]
−1 −1 2
1 1 −3 −1 4
Next, 𝐴−1
𝐿 = det(𝐴) 𝑎𝑑𝑗(𝐴) = 2 [ ]
−1 −1 2
1 −3 −1 4 −2 1 2 1
𝒙 = 𝐴−1
𝐿 𝒃= [ ] [ 4 ]= [ ]=[ ]
2 −1 −1 2 2 −2 −1
0
92
We note that in this example, adjoint method and pseudo method give the same
solution (see example 6.1.2).
2𝑥1 + 3 𝑥2 − 2𝑥3 = 4
−6𝑥1 − 8 𝑥2 + 6𝑥3 = 1
2 3 −2 4
Hold by 𝐴 = [ ] is the coefficient matrix, and 𝒃 = [ ] is the constant
−6 −8 6 1
1
vector. 𝑑𝑒𝑡(𝐴) = 4, since 𝐴 is 2 × 3, 𝐴−1
𝑅 exists, 𝒙 = 𝑎𝑑𝑗(𝐴) 𝒃
det(𝐴)
−14 −5
But, 𝑎𝑑𝑗(𝐴) = [ 12 4]
2 1
−14 −5
1 1
Next, 𝐴−1
𝑅 = 𝑎𝑑𝑗(𝐴) = [ 12 4]
det(𝐴) 4
2 1
1 1 −14 −5 4 1 −61
𝒙= 𝑎𝑑𝑗(𝐴) 𝒃 = [ 12 4 ] [ ] = [ 52 ]
det(𝐴) 4 1 4
2 1 9
−71 52 9
That is, x = ( , , )
4 4 4
Here the solution we have obtained using adjoint method is different from the
pseudo method solution (see example 6.1.5) and both of them are members from
−35
4 1 0 0 −16 −6 𝑡1
1 2 3 −2
𝑥 = [ 13 ] + ([0 1 0] − 8 [ 24 8 ][ ]) [𝑡2 ]
35 −6 −8 6 𝑡3
0 0 1 16 6
4
−35
4 1
1 −71 52 9
𝑥 = [ 13 ] + 2 𝑡 [ 0], Choose 𝑡 = −13, x = ( , , ).
35 4 4 4
1
4
93
6.3 Cramer's rule for nonsingular 𝒎 × 𝒏 matrices
In linear algebra, Cramer's rule (see theorem 1.7.4) gives an explicit formula for the
solution of a system of linear equations with as many equations as unknowns. That is, for
the solution of a system with a square matrix and provided that the coefficient matrix is
invertible, Cramer’s rule offers a simple and a convenient formula for the solution. In this
section we want to generalize this method for an 𝑚 < 𝑛 system of linear equations. As in
the usual method of Cramer's, the result for rectangular matrices uses the minors of a
matrix. We also use the results in order to solve a matrix equation. in the case of systems
with an infinite number of solutions, get the final formula for calculating the unknowns by
The key to Cramer’s Rule is replacing the variable column of interest with the
constant column and calculating the determinants.
𝑏1
𝑏
𝐴 = [𝑎𝑖𝑗 ] be the coefficient matrix, 𝑏 = [ 2 ] . If det(𝐴) ≠ 0. Then
⋮
𝑏𝑚
94
proof: We look at the linear system 𝑨𝒙 = 𝒃.
𝒙 = 𝐴−1 𝒃
𝑥1
𝑥2 1
[ ⋮ ] = det(𝐴) (𝑎𝑑𝑗 𝐴) 𝒃
𝑥𝑛
𝑥1
𝐴11 𝐴21 ⋯ 𝐴𝑚1
𝑥2 1
[ ⋮ ] = det(𝐴) [ ⋮ ⋮ ⋱ ⋮ ] 𝒃
𝑥𝑛 𝐴1𝑛 𝐴2𝑛 ⋯ 𝐴𝑚𝑛
Where
Hence
det(𝐴𝑗 )
𝑥𝑗 = .
det(𝐴)
95
For 𝑗 = 1, 2, … , 𝑛. In this expression for 𝑥𝑗 , the determinant of 𝐴𝑗 , det(𝐴𝑗 ), can be
calculated by any method. It was only in the derivation of the expression for 𝑥𝑗 that we
2𝑥1 + 3 𝑥2 − 2𝑥3 = 4
−6𝑥1 − 8 𝑥2 + 6𝑥3 = 1
2 3 −2 4
Here the coefficient matrix is 𝐴=[ ], 𝑑𝑒𝑡(𝐴) = 4, and 𝒃 = [ ] is the
−6 −8 6 1
constant vector.
4 3 −2 det(𝐴1 )
𝐴1 = [ ] , det(𝐴1 ) = −59 𝑥1 = = −59⁄4
1 −8 6 det(𝐴)
2 4 −2 det(𝐴2 )
𝐴2 = [ ] , det(𝐴2 ) = 52 𝑥2 = = 52⁄4
−6 1 6 det(𝐴)
2 3 4 det(𝐴3 )
𝐴3 = [ ] , det(𝐴3 ) = 11 𝑥3 = = 11⁄4
−6 −8 1 det(𝐴)
We note that the general solution for this example given by Theorem 6.1.6 is
−35
4 1 1
𝑥 = 13 + 𝑡 [ 0]
35 2
1
[ 4 ]
96
Theorem 6.3.3 [10 ]. (Generalization of Cramer's rule )
Also, this solution coincides with the Cramer's rule formula when 𝑛 = 𝑚. In
𝑚
𝑑𝑒𝑡((𝐴 𝐴𝑇 ))𝑖
𝑥𝑗 = ∑ 𝑎𝑖𝑗 , 𝑗 = 1, 2, 3 , … , 𝑛,
𝑑𝑒𝑡(𝐴 𝐴𝑇 )
𝑖=1
𝑏1
𝑏
𝐴 𝐴𝑇 by the entries in the matrix [ 2 ]
⋮
𝑏𝑚
The system 𝑨𝒙 = 𝒃 is solvable for all 𝑏 ∈ 𝑅 𝑚 , if and only if, the operator 𝐴 is
onto (see theorem 1.9.15).
i.e, 𝑅𝑎𝑛𝑔𝑒(𝐴) = 𝑅 𝑛 Hence, from the lemma 1.8.15 there exists 𝛾 > 0 such that
‖𝐴𝑇 𝑧‖𝑅𝑚 ≥ 𝛾 ‖𝑧‖𝑅𝑛 , 𝑧 ∈ 𝑅 𝑛 .
Therefore,
〈𝐴 𝐴𝑇 𝑧 , 𝑧〉 ≥ 𝛾 2 ‖𝑧‖2 𝑅𝑛 , 𝑧 ∈ 𝑅 𝑛 .
97
Suppose now that 𝑑𝑒𝑡(𝐴 𝐴𝑇 ) ≠ 0. Then (𝐴 𝐴𝑇 )−1 exists and given 𝑏 ∈ 𝑅𝑛 we
𝑑𝑒𝑡((𝐴 𝐴𝑇 )𝑗 )
det((𝐴 𝐴𝑇 )1 )
∑𝑛𝑗=1 𝑎𝑗,1
det(𝐴 𝐴𝑇 ) 𝑑𝑒𝑡(𝐴 𝐴𝑇 )
𝑎11 𝑎21 ⋯ 𝑎𝑛1
det((𝐴 𝐴𝑇 )2 ) 𝑑𝑒𝑡((𝐴 𝐴𝑇 )𝑗 )
𝑎12 𝑎22 ⋯ 𝑎𝑛2 ∑𝑛𝑗=1 𝑎𝑗,2
𝑥𝑗 = [ ⋮ ⋱ ⋮ ]
⋮ det(𝐴 𝐴𝑇 ) = 𝑑𝑒𝑡(𝐴 𝐴𝑇 )
𝑎1𝑚 𝑎21 ⋮ ⋮
… 𝑎𝑛𝑚 det((𝐴 𝐴𝑇 )𝑚 )
𝑑𝑒𝑡((𝐴 𝐴𝑇 )𝑗 )
[ det(𝐴 𝐴𝑇 ) ] 𝑛
[∑𝑗=1 𝑎𝑗,𝑚 𝑑𝑒𝑡(𝐴 𝐴𝑇 ) ]
2𝑥1 + 3 𝑥2 − 2𝑥3 = 4
−6𝑥1 − 8 𝑥2 + 6𝑥3 = 1
2 3 −2 4
Here the coefficient matrix is 𝐴=[ ], 𝑑𝑒𝑡(𝐴) = 4, and 𝑏 = [ ] is the
−6 −8 6 1
constant vector.
2 −6
𝑇 2 3 −2 17 −48
𝐴𝐴 =[ ] [ 3 −8] = [ ] → 𝑑𝑒𝑡(𝐴 𝐴𝑇 ) = 8
−6 −8 6 −48 136
−2 6
98
4 −48
(𝐴 𝐴𝑇 )1 = [ ] → 𝑑𝑒𝑡(𝐴 𝐴𝑇 )1 = 592
1 136
(𝐴 𝐴𝑇 )2 = [ 17 4
] → 𝑑𝑒𝑡(𝐴 𝐴𝑇 )2 = 209
−48 1
We find that the solution of the system in this way is the same as pseudo
solution given in example 6.2.2 which comes from the general solution (see
theorem 6.1.6 )
−35
4 1
1
𝑥 = [ 13 ] + 2 𝑡 [ 0], at 𝑡=0
35
1
4
That is both a Generalization of Cramer's rule and pseudo method give the
same solution.
illustrate the results of what we have done in the previous sections especially in
Example 6.4.1 [10 ]. Consider the following particular case of the system 𝑨𝒙 = 𝒃
99
𝑎11
𝑎12
Then, if we define the column vector 𝐼1 = [ ⋮ ],
𝑎1𝑛
𝑎11
𝐴 𝐴𝑇 = [𝑎11 𝑎12 … 𝑎1𝑛 ] [𝑎12 ] = ‖𝐼1 ‖2 . (see definition 1.8.2)
⋮
𝑎1𝑛
1
Then, (𝐴 𝐴𝑇 )−1 𝒃 = ‖𝐼1 ‖2
𝒃 and
𝑎1𝑗 𝑏 𝑎1𝑗 𝑏
𝑥𝑗 = 2
= 𝑛 2 , 𝑗 = 1,2, … , 𝑛
‖𝐼1 ‖ ∑𝑗=1 𝑎1𝑗
𝑎11 𝑎21
𝑎12 𝑎22
𝐼1 = [ ⋮ ], 𝐼2 = [ ⋮ ].
𝑎1𝑛 𝑎2𝑛
Then,
𝑎11 𝑎21
𝑎11 𝑎12 ⋯ 𝑎1𝑛 𝑎12 𝑎22 ‖𝐼1 ‖2 〈𝐼1 , 𝐼2 〉
𝐴 𝐴𝑇 = [𝑎 𝑎22 ⋯ 𝑎2𝑛 ] [ ⋮ ⋮ ] = [ ].
21 〈𝐼2 , 𝐼1 〉 ‖𝐼2 ‖2
𝑎1𝑛 𝑎2𝑛
100
1 ‖𝐼2 ‖2 −〈𝐼1 , 𝐼2 〉
(𝐴 𝐴𝑇 )−1 = [ ].
‖𝐼1 ‖2 ‖𝐼2 ‖2 − |〈𝐼1 , 𝐼2 〉|2 −〈𝐼2 , 𝐼1 〉 ‖𝐼1 ‖2
𝑥1 𝑎11 𝑎21
𝑥2 1 𝑎12 𝑎22 ‖𝐼2 ‖2 −〈𝐼1 , 𝐼2 〉 𝑏1
[ ⋮]= [ ] [ ] [ ].
‖𝐼1 ‖2 ‖𝐼2 ‖2 − |〈𝐼1 , 𝐼2 〉|2 ⋮ ⋮ −〈𝐼2 , 𝐼1 〉 ‖𝐼1 ‖2 𝑏2
𝑥𝑛 𝑎1𝑛 𝑎2𝑛
𝑥1 + 𝑥2 = 1
−𝑥1 + 𝑥2 + 𝑥3 = −1
1 −1
1
𝐼1 = [1] , 𝐼2 = [ 1 ] , 𝑏=[ ]
−1
0 1
101
𝑏1 ‖𝐼2 ‖2 − 𝑏2 〈𝐼1 , 𝐼2 〉 𝑏2 ‖𝐼1 ‖2 − 𝑏1 〈𝐼2 , 𝐼1 〉
𝑥1 = 𝑎11 + 𝑎 21
‖𝐼1 ‖2 ‖𝐼2 ‖2 − |〈𝐼1 , 𝐼2 〉|2 ‖𝐼1 ‖2 ‖𝐼2 ‖2 − |〈𝐼1 , 𝐼2 〉|2
1(1 × 3 − −1 × 0) (−1)(−1 × 2 − 1 × 0) 3 2 5
= + = + =
6 6 6 6 6
1(3 − 0) (1)(−2 − 0) 3 −2 1
= + = + =
6 6 6 6 6
(0)(3) (1)(−2) −2
= + =
6 6 6
We note that if ‖𝐼1 ‖2 ‖𝐼2 ‖2 = |〈𝐼1 , 𝐼2 〉|2 this means that the angle between
𝐼1 , 𝐼2 the equals zero and this indicates that the system of equations are identical,
meaning that there is an infinite number of solutions and in this case it is the
But if the angle between 𝐼1 , 𝐼2 is not equal to zero, then this means that the
We note that the general solution for this example given by Theorem 6.1.6 is
1 5 1 1
𝑥 = [ 1 ] + 𝑡 [ −1]
6 6
−2 2
102
Example 6.4.4 [10 ]. Consider the following general case of system 𝑨𝒙 = 𝒃
‖𝐼1 ‖2 0 0 ⋯ 0
‖𝐼2 ‖2 ⋯ 0
𝐴 𝐴𝑇 = [ 0 0
⋮ ]
⋮ ⋮ ⋮ ⋱
0 0 0 ⋯ ‖𝐼𝑛 ‖2
−𝑥1 − 𝑥2 + 𝑥3 + 𝑥4 = 1
−𝑥1 + 𝑥2 − 𝑥3 + 𝑥4 = 1
𝑥1 − 𝑥2 − 𝑥3 + 𝑥4 = 1
Solution:
−1 −1 1
1
𝐼1 = [−1] , 𝐼2 = [ 1 ], 𝐼3 = [−1] , 𝑏 = [1]
1 −1 −1
1
1 1 1
103
𝑎13 𝑏1 𝑎23 𝑏2 𝑎33 𝑏3 (1)(1) (−1)(1) (−1)(1) −1
𝑥3 = + + = + + =
‖𝐼1 ‖2 ‖𝐼2 ‖2 ‖𝐼3 ‖2 4 4 4 4
We note that the general solution for this example given by Theorem 6.1.6 is
−1 1
1 −1 1 1]
𝑥 = [ ]+ 𝑡[
4 −1 4 1
3 1
−1 −1 −1 3
Choose 𝑡 = 0, 𝑥=( , , , )
4 4 4 4
104
References: -
[1] Amiri, M., Fathy, M., Bayat, M., Generalization of some determinantal identities for
non-square matrices based on Radics definition, TWMS J. Pure Appl. Math. 1, no 2(2010),
163-175.
[2] Anton, H., Ilementary Linear Algebra with application, Seventh edition.
[3] Arunkumar, M., Murthy, S., and Ganapath, G., Determinant for non-square
matrices, IJMSEA, no 5 (2011), 389-401.
[7] Joshi, V.N., A determinant for rectangular matrices, no. 21 (1980) , 137-146.
[9] Kolman, B., Introductory Linear Algebra with application, Six edition.
105
[13] Radic, M.,A bout a determinant of rectangular 2 × 𝑛 matrix and its
geometric interpretation, Beitrage Algebra Geom. 46,no.1(2005),321-349.
[14] Radić, M., A definition of determinant of rectangular matrix, Glas. Mat. Ser.
III 1(21) (1966), 17–22.
106