1bis-Linear Algebra Lectures Notes NYU
1bis-Linear Algebra Lectures Notes NYU
Pyxis
20
10
K10
K20
K10 0 10 20
> display en , a n , b n , dn , Ordo, Abs , color = "LightBlue", axes = boxed, scaling = constrained, title
= `Pyxis` ;
L����� A������
Math -UA 9140
L������ N����
Joachim L�������
S����� 2023
I Matrices Calculus 1
1 Matrices Calculus 2
1.1 De�nition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 Notations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Particular Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.4 Operations on Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
First Operation: Transpose of a Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
Second Operation: Multiplication of a Matrix by a Scalar . . . . . . . . . . . . . . . . . 4
Third Operation: Addition of two Matrices . . . . . . . . . . . . . . . . . . . . . . . . . 5
Fourth Operation: Multiplication of two Matrices . . . . . . . . . . . . . . . . . . . . . 5
1.5 The ring of Square Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.6 Invertible Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2 Linear Systems 12
2.1 Linear System as a Matrix Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.2 Elementary Operations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
2.3 Gaussian Elimination & Row Reduced Echelon Matrices . . . . . . . . . . . . . . . . . . 16
2.4 Application of Gaussian Elimination to Inversion of Matrices . . . . . . . . . . . . . . . 21
2.5 Application of Gaussian Elimination to LU Factorization . . . . . . . . . . . . . . . . . 23
LU Factorization of Regular Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
P A = LU Factorization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.6 Application of Gaussian Elimination to solving Linear System . . . . . . . . . . . . . . 31
Results for general linear systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
Results About Square Linear Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
3 Vector Spaces 41
3.1 Vector Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
De�nition & Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.2 Vector Subpaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
De�nition & Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
3.3 Intersection and Sum of Vector Subspaces . . . . . . . . . . . . . . . . . . . . . . . . . . 45
Intersection of Vector subspaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
Sum of Vector subspaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
Supplementary Vector Subspaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
3.4 Spanning Family & Linearly Independent Family of Vectors . . . . . . . . . . . . . . . . 49
ii
L����� A������ Lecture Notes C�������
5 Use of Basis 70
5.1 Presentation of the problem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
5.2 Matrix Representation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
Vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
Linear Maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
Fondamental Formulae . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
A Particular kind of linear maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
Rank of a Linear Map . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
5.3 Change of Basis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
Notations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
The case of vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
The case of Linear Maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
The case of Endomorphisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
IV Diagonalization 89
7 Diagonalization 90
7.1 Eigenvalues of an Endomorphism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
De�nitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
Applications to the Calculation of Eigenvalues . . . . . . . . . . . . . . . . . . . . . . . 92
7.2 Diagonalizable Endomorphisms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
7.3 Diagonalizable Matrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
De�nition and Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
Diagonalization in practice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
7.4 Applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
V Multilinear Algebra 99
8 Isometries 100
VI Applications 102
Bibliography 111
Index 112
iv S����� 2023
Greek Letters and their correspondance
in the Latin alphabet
Greek letters
Latin letters Classical use in Math
Lowercase letter Capital letter Name
– A alpha a
— B beta b
“ gamma g
” delta d
Á E epsilon e Á: in�nity small
’ Z zeta z
÷ H eta e
◊ theta th ◊: angle
ÿ I iota i
Ÿ K kappa k
⁄ lambda l
µ M mu m
‹ N nu n
› xi x
o O omicron o
r universal cst: fi ¥ 3.141592653589793
fi pi p r
: product
fl P rho r
q q
‡ sigma s : Sum
· T tau tau
‚ Y upsilon u
„, Ï phi f
‰ X khi kh
 psi ps
Ê omega o
v
Part I
1
C������ 1
Matrices Calculus
Throughout this course K denotes either the set of real numbers, denoted R or the set of complex
numbers, denoted C. Moreover, p and q will denote two positive integers.
1.1 De�nition
We denote M(p, q, K) or Mp,q (K) the set of all p ◊ q matrices with coe�cients in K. In the case
where p = q, we use the shorthand notation Mp (K).
1.2 Notations
In order to represent a p ◊ q matrix A, we usually use the following notation
à í
a1,1 a1,2 · · · a1,q
a2,1 a2,2 · · · a2,q
A := .. .. .. .
. . .
ap,1 ap,2 · · · ap,q
Remark 1.2.1
It is important to respect the convention on the order of indices: we denote ai,j the coe�cient of the
matrix A which is on the ith row and j th column.
When one wants to give the coe�cients of A with a formula, one can also use a condensed notation:
2
L����� A������ Lecture Notes C������ 1: Matrices Calculus
Example 1.2.2
Denote A := (i ≠ j)16i62,16j63 . It is clear that:
Ç å
0 ≠1 ≠2
(i ≠ j)16i62,16j63 = .
1 0 ≠1
When there is no ambiguity because we do know the “dimensions” p and q of the matrix A, we simply
write A := (ai,j ). In particular, we will denote (0), and one will call null matrix, the matrix �r which
all coe�cients equal 0. Hence there is a null matrix for every couple of integers (p, q). They are all
denoted the same way. It does not provoke any confusion unless we do not take care to the context.
A = (a1 , a2 , · · · , aq ).
3 S����� 2023
L����� A������ Lecture Notes C������ 1: Matrices Calculus
Example 1.4.1
De�ne the following matrices
Ç å Ç å
2 ≠2 3 ≠1 2 3
A := & B := .
4 3 7 9 ≠5 6
It is clear that A ”= B since a1,1 ”= b1,1 .
Example 1.4.3
Ö è
2 4
The transpose of the matrix A, de�ned in Example 1.4.1, is: ≠2 3 .
3 7
(⁄ · A)i,j := ⁄ · ai,j .
Example 1.4.5
If we still consider Matrix A, de�ned at Example 1.4.1, we have:
Ç å Ç å
2 ≠2 3 4 ≠4 6
2·A=2· = .
4 3 7 8 6 14
4 S����� 2023
L����� A������ Lecture Notes C������ 1: Matrices Calculus
Let ⁄ be a real number. We have t(⁄ · A) = ⁄ · tA. Moreover, one can have ⁄ · A = (0) if and only if
⁄ = 0 or if A = (0).
Example 1.4.7
Having in mind matrices A and B, de�ned at Example 1.4.1, we can write:
Ç å Ç å Ç å Ç å
2 ≠2 3 ≠1 2 3 2≠1 ≠2 + 2 3 + 3 1 0 6
A+B = + = = .
4 3 7 9 ≠5 6 4 + 9 3≠5 7 + 6 13 ≠2 13
For ay real numbers ⁄ and µ, it is easy to see that the following rules hold.
(⁄ + µ) · A = ⁄ · A + µ · A, ⁄ · (A + B) = ⁄ · A + ⁄ · B, (A + B) = tA + tB.
t
(4.1)
The readers who are familiar with Group Theory noticed that (Mp,q (K), +) is an abelian group.
Example 1.4.9
We hence have:
Ö è
Ç å ≠1 9 Ç å
2 ≠2 3 2 · (≠1) + (≠2) · 2 + 3 · 3 2 · 9 + (≠2) · (≠5) + 3 · 6
· 2 ≠5 =
4 3 7 4 · (≠1) + 3 · 2 + 7 · 3 4 · 9 + 3 · (≠5) + 7 · 6
3 6
Ç å Ç å
≠2 ≠ 4 + 9 18 + 10 + 18 3 46
= = .
≠4 + 6 + 21 36 ≠ 15 + 42 23 63
Remark 1.4.10
In order to de�ne the product of a matrix A by a matrix B, we need that the number of columns of
A equals the number of rows of B. The matrix AB has as much rows as A and as much columns as
B.
5 S����� 2023
L����� A������ Lecture Notes C������ 1: Matrices Calculus
Example 1.4.11
Let x1 , x2 , y1 and y2 be real numbers. Let’s compute the two following products.
Ç å Ç å Ç å
y1 y1 y1 x1 y1 x2
(x1 , x2 ) · = (x1 y1 + x2 y2 ), and · (x1 , x2 ) = .
y2 y2 y2 x1 y2 x2
Let ⁄ be a real number. It is easy to see that the following rule hold.
For the remainder of this section denote r and s two positive integers.
Proposition 1.4.12
The product of two matrices is an associative law. More precisely, if A is a matrix in Mp,q (K), B is
a matrix in Mq,r (K) and C is a matrix in Mr,s (K) then:
(AB) C = A (BC).
Proof 1.4.13
We know that AB belongs to Mp,r (K), and then (AB) C belongs to Mp,s (K). We also know that
BC belongs to Mq,s (K) and then A (BC) belongs to Mp,s (K). Let us show that, for all (i, j) in
J1, pK ◊ J1, sK:
((AB) · C)i,j = (A · (BC))i,j
For all (i, j) in J1, pK ◊ J1, sK, one can write:
r r
Ç q å q
r ÿ
ÿ ÿ ÿ ÿ
((AB) · C)i,j = (AB)i,k ck,j = ai,¸ b¸,k ck,j = ai,¸ b¸,k ck,j
k=1 k=1 ¸=1 k=1 ¸=1
and
q q Ç r
å q ÿ
r q
r ÿ
ÿ ÿ ÿ ÿ ÿ
(A · (BC))i,j = ai,¸ (BC)¸,j = ai,¸ b¸,k ck,j = ai,¸ b¸,k ck,j = ai,¸ b¸,k ck,j ,
¸=1 ¸=1 k=1 ¸=1 k=1 k=1 ¸=1
Proposition 1.4.14
For any matrices A in Mp,q (K) and B in Mq,r (K), we have:
t
(AB) = tB tA.
Proof 1.4.15
Let A := (ai,j )(i,j)œJ1,pK◊J1,qK and B := (bi,j )(i,j)œJ1,qK◊J1,rK , one can write, for all (i, j) in J1, pK ◊
6 S����� 2023
L����� A������ Lecture Notes C������ 1: Matrices Calculus
J1, rK,
q
ÿ
( (AB))i,j = (AB)j,i =
t
aj,k bk,i
k=1
&
q
ÿ q
ÿ q
ÿ
( B A)i,j =
t t t
B i,k
t
A k,j
= bk,i aj,k = aj,k bk,i ,
k=1 k=1 k=1
Proposition 1.4.16
The product of matrices is distributive over addition. More precisely, if A and B are matrices of
Mp,q (K) and C and D are matrices of Mq,r (K), then
A · (C + D) = A · C + A · D and (A + B) · C = A · C + B · C.
Proof 1.4.17
It is clear that A · (C + D), A · C and A · D, (A + B) · C and B · C all belong to Mp,r (K). Moreover,
for all (i, j) œ J1, pK ◊ J1, rK, one can write, from one hand:
n
ÿ n
ÿ
(A · (C + D))i,j = ai,k (C + D)k,j = ai,k (ck,j + dk,j )
k=1 k=1
and
n
ÿ n
ÿ n
ÿ
(A · C + A · D))i,j = (A · C)i,j + (A · D)i,j = ai,k ck,j + ai,k dk,j = ai,k (ck,j + dk,j )
k=1 k=1 k=1
and
n
ÿ n
ÿ n
ÿ
(A · C + B · C)i,j = (A · C)i,j + (B · C)i,j = ai,k ck,j + bi,k ck,j = (ai,k + bi,k ) ck,j ,
k=1 k=1 k=1
7 S����� 2023
L����� A������ Lecture Notes C������ 1: Matrices Calculus
Example 1.4.19
De�ne the following matrices:
Ö è
Ç å 3 9 ≠9
0 1
C := , D := 2 0 0 .
0 0
3 3 ≠3
where ”i,j denotes the Kronecker symbol which value is 1 if i = j and 0 otherwise. One can easily
check that:
Ip · A = A · Ip = A,
for every p ◊ p matrix A.
1
A ring is a set R equipped with two binary operations, denoted + and ·, satisfying the following three sets of axioms,
called the ring axioms:
1. (R, +) is an abelian group. In other words:
8 S����� 2023
L����� A������ Lecture Notes C������ 1: Matrices Calculus
Remark 1.5.1
1. Note that (Mp (K), ·) is not commutative if p > 1. Hence we have:
Ç å Ç å Ç å Ç å Ç å Ç å
1 0 1 1 1 1 1 1 1 0 1 2
· = and · = .
0 2 0 1 0 2 0 1 0 2 0 2
2. Moreover (Mp (K)+, ·) is not an integral domain. That means that one can �nd two non
zero matrices such that their product equals the null matrix. Hence:
Ç å Ç å Ç å
0 1 0 1 0 0
· = = 0M2 (K) .
0 0 0 0 0 0
Let’s �nish this section by a result on multiplication of matrices, the proof of which can be performed
by induction, on the integer n.
Remark 1.5.3
One can apply the previous result to compute An , once A as been expanded as a sum of a diagonal
matrix D and a nilpotenta matrix N .
a
i.e. there exists a positive integer q such that N q = 0.
Proof 1.5.4
This is nothing but the Binomial Formula, which applies when the two elements commute. ⇤
B · A = A · B = Ip .
The matrix B is denoted A≠1 and is called invert of the matrix A. We also say in this case that A is
a regular matrix.
9 S����� 2023
L����� A������ Lecture Notes C������ 1: Matrices Calculus
Remark 1.6.2
A matrix can be invertible only if it is a square matrix.
Denote GLp (K) the set of all invertible matrices of Mp (K). It is called the General linear group of
degree p.
Example 1.6.3
Ç å
1 2
1. It is easy to verify that if A = then, the inverse of A, denoted A≠1 , is A≠1 =
3 4
Ç å Ç å
≠2 1 1 2
3 1
. Besides, if B = 4
then, the inverse of B, denoted B ≠1 , is B ≠1 =
≠ 3
Ç 22 2
5
å 5
≠ 13 13
15 5
.
26 ≠ 26
Ç å
a b
2. De�ne A the element of M2 (K) by setting: A := . We will show, during recitations,
c d
that A is invertible if and only if ad ≠ bc ”= 0. When this is the case, we have the equality:
Ç å
1 d ≠b
A ≠1
= · .
ad ≠ bc ≠c a
Remark 1.6.4
1. If A is invertible and if G · A = I, then G = GAA≠1 = A≠1 . Likewise, if AD = I, then
D = A≠1 . In particular, the inverse of a matrix, when it exists, is unique.
2. If there are two matrices G and D such that GA = AD = I, then we have G = GAD = D
and A is invertible and its inverse is A≠1 = G = D.
3. We will see later that, if there exists a matrix G such that GA = I, then A is an invertible
matrix and A≠1 = G.
Proposition 1.6.5
Let A and B be two matrices in GLp (K) then:
1. AB is an invertible matrix and its inverse is B ≠1 A≠1 . In other words, the inverse of a product
of matrices is the product of the inverses, but in reverse order.
Proof 1.6.6
1. The fact that AB is invertible comes from the fact that AB · B ≠1 A≠1 = A · A≠1 = Ip .
Theunicity of an inverse, when it exists leaves no other choice but B ≠1 A≠1 for the inverse of
AB.
10 S����� 2023
L����� A������ Lecture Notes C������ 1: Matrices Calculus
2. It is su�cient to show that tA · t A≠1 = Ip . For all (i, j) in J1, pK2 , one can write:
p
ÿ p
ÿ p
ÿ
A · t A≠1
t
i,j
= (tA)i,k (tA≠1 )k,j = ak,i (A≠1 )j,k = (A≠1 )j,k Ak,i = A≠1 · A j,i
k=1 k=1 k=1
= (Ip )i,j .
This latter equality shows that tA · t A≠1 = Ip and achieves the proof.
Example 1.6.7
1. Let A and B be the two matrices de�ned at Example 1.6.3. It is easy to check that:
Ç 18
å Ç 23 9
å
7 5 26 ≠ 26
(A · B) ≠1
= 46
≠1
=
15 5 ≠ 75
52
35
52
while Ç 2 å Ç å Ç å
5 23 9
≠ 13 13 ≠2 1 26 ≠ 26
B ≠1
·A ≠1
= 15 5
· 3 = .
26 ≠ 26 2 ≠ 12 ≠ 75
52
35
52
2. Besides,
Ç å≠1 Ç å
1 3 ≠2 32
(t A)≠1 = =
2 4 1 ≠ 12
&
Ç å Ç å
≠2 1 ≠2 32
t
(A ≠1
)= t
3
= .
2 ≠ 12 1 ≠ 12
11 S����� 2023
C������ 2
Linear Systems
where aij and bj are elements of K and the xi are the unknowns that we want to determine. A solution
of the system is a q-tuple denoted (x1 , x2 , · · · , xq ) of Kq for which the p equalities are satis�ed. We
are now interested to the the set of all solutions of the system. Let’s introduce the following notations:
à í à í
x1 b1
x2 b2
A := (ai,j ), X := .. , & B := .. .
. .
xq bp
(SÕ ) AX = B.
Matrix A is called matrix of the system and the column matrix B is called second member of the system.
When B = (0) (i.e. bi = 0, for all 1 6 i 6 p), the system is said to be homogeneous. In this case it has
an obvious solution that is the null solution (0, · · · , 0). This latter is also called trivial solution.
Remark 2.1.2
Two equivalent linear systems have the same number of unknowns but they might have di�erent
12
L����� A������ Lecture Notes C������ 2: Linear Systems
are equivalent.
Proposition 2.1.3
Let A be a p◊q matrix and let G be a p◊p invertible matrix. The systems AX = B and GAX = GB
are equivalent.
Proof 2.1.4
The proof is easy. Indeed, assuming AX = B holds, we just have to multiply, from the left hand
side, both sides of the previous equality by G to get GAX = GB.
Conversely, if GAX = GB holds, we just have to multiply, from the left hand side, both sides of the
previous equality by G≠1 to get AX = B. ⇤
Remark 2.1.5
In the particular case where A is an invertible matrix (i.e. A belongs to in GLp (K)),
then this system is equivalent to the system X = A≠1 B, which has a unique solution
((A≠1 B)1 , (A≠1 B)2 , · · · , (A≠1 B)p ). Such a system is called a Cramer (or regular) system.
More precisely, denote S(i, j) the operation which consists of adding the j th equation to the ith one.
Note that this means exactly adding the j th row of A to the ith one.
As we will see each of these operations is nothing but multiplying the matrix A of the system (SÕ )
by an invertible matrix, which is said to be associated to the elementary operation. As we will notice
further, the matrix associated to any of the elementary operations described above can be obtained by
applying this elementary operation to the unit matrix.
13 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
Let A be a matrix with p rows. We easily see that the matrix M (⁄, i) · A has the same rows as A except
for the ith row, for which all coe�cients has been multiplied by ⁄.
This means that, applying the operation M(⁄, i) to the system AX = B gives us the system
M (⁄, i) · AX = M (⁄, i) · B.
It is also easy to see that the operation M( ⁄1 , i) "cancels" the e�ect of M(⁄, i). The consequence of
this fact is that Å ã
1
M (⁄, i) · M , i = I.
⁄
In other words, for all ⁄ in Rú , M (⁄, i) is invertible and:
Å ã
1
M (⁄, i) = M
≠1
,i .
⁄
Let A be a matrix with p rows. We easily see that the matrix P (i, j) · A has the same rows as A except
for that the rows i and j have been interchanged.
14 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
This means that, applying the operation P(i, j) to the system AX = B gives us the system
P (i, j) · AX = P (i, j) · B.
It is also easy to see that the operation P(j, i) "cancels" the e�ect of P(i, j). The consequence of
this fact is that
P (i, j) · P (j, i) = I.
In other words, for all (i, j) in J1, pK2 , P (i, j) is invertible and:
Let A be a matrix with p rows. We easily see that the matrix S(i, j) · A is obtained from A by replacing
its j th row by the sum of the ith and j th rows. applying operation S(i, j) to the system AX = B
gives us the system
S(i, j) · AX = S(i, j) · B.
"Canceling" operation S(i, j) can be done by subtracting the ith row of the new matrix (i.e. S(i, j) · A)
to its j th row. More precisely we can do that in three steps.
In other words, for all (i, j) in J1, pK2 , the matrix S(i, j) is invertible and:
15 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
One can generalize the three previous operations. Indeed, for any non zero real ⁄,
Å ã
1
S(⁄, i, j) := M , i · S(j, i) · M (⁄, i).
⁄
corresponds to the operation:“adding to the j th row the ith row multiplied by ⁄”. We denote S(⁄, i, j)
this operation.
Remark 2.2.1
Note that the matrices of these successive operations are written from the right to the left since,
every time, one multiplies from the left hand side.
One easily sees that S(⁄, i, j) is an invertible matrix. Moreover, one has:
Å ã
1
S(⁄, i, j) := M ≠ , i · S(j, i) · M (≠⁄, i) = S(≠⁄, i, j).
≠1
⁄
The entries indicated by ~ are the pivots, and must be nonzero. The �rst r rows of R each contain
exactly one pivot, but not all columns are required to include a pivot entry. The entries below the
“staircase”, indicated by the solid line, are all zero, while the non-pivot entries above the staircase,
indicated by stars, can be anything. The last p ≠ r rows are identically zero and do not contain any
pivots. there may, in exceptional situations, be one or more all zero initial columns.
Example 2.3.2
The matrix â ì
4 1 0 ≠4 15 9
0 ≠2 1 6 7 ≠2
R :=
0 0 0 0 ≠3 ≠1
0 0 0 0 0 0
is in row echelon form. It has three pivots that are: 4, ≠2 and ≠3.
Note moreover that the row echelon matrix can have several initial columns consisting of all zeros.
16 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
This matrix is also in row echelon form. It has three pivots that are: 2, ≠3 and ≠9.
Example 2.3.4
The matrix 0 1
0 0 1 ú ú ú 0 ú 0 ú ú
B C
B0 0 0 0 0 0 1 ú 0 ú úC
B C
T := B
B0 0 0 0 0 0 0 0 1 ú úCC
B C
@0 0 0 0 0 0 0 0 0 0 0A
0 0 0 0 0 0 0 0 0 0 0
is in reduced row echelon form. Moreover, the pivot columns are here highlighted in bold font.
Example 2.3.6
The rank of Matrix T , given at Example 2.3.4, is 3.
17 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
Proposition 2.3.7
The rank of a reduced row echelon matrix is its number of pivots.
Remark 2.3.8
In the particular case where R is reduced row echelon matrix with rank 0, we see that one can only
add columns full of zeros. Thus all the coe�cients of R are zeros. Thus R = (0).
If it is easy to determine the rank of a reduced row echelon matrix. However, one needs a concrete
way to determine the rank of any matrix. This can be done thanks to the following result, the proof of
which gives a concrete way to do so.
GA = R.
Let us now prove Theorem 2.3.9, the proof of which is known as Gaussian Elimination.
Proof 2.3.11
Unicity:
The unicity of such a matrix G, assuming it exists, is a straight consequence of the unicity of the
inverse of a given matrix (see 1.6.4).
Existence of G:
The Matrix G will be built as a product of matrices associated to the elementary operations described
in Section 2.2. In view of Proposition 2.1.3, Matrix G will be invertible since it is a product of invertible
matrices.
We therefore just have to show that, by a succession of elementary operations, one can transform
Matrix A := (ai,j )(i,j)œJ1,pK◊J1,qK into a reduced row echelon matrix.
Assume that the the �rst column of A has at least one non zero coe�cient, say ak,1 . By multiplying
row k by a1k1 we now are in the case where the �rst column of A has one coe�cient, namely ak1 ,
equals to 1 exactly. By interchanging rows 1 and k, one can reduce our reasoning to the case where
this coe�cient (equal to 1) is on the �rst row of A.
If we add successively, for i from 2 to p, the �rst row multiplied by ≠ai1 to the ith row, we get a
matrix the �rst column of which is full of zeros except for the �rst element which equals 1. One can
18 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
In order to prove the theorem we do an induction on p. The �rst part of the demonstration proved
that, by elementary operations on the rows of A, can get a matrix the �rst column of which is of the
form: à í
Ü ê 1
0
.. 0
. or .. .
.
0
0
• If q = 1 (or if p = 1)
19 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
then this ends the proof and the matrix we obtain is in reduced row echelon form. Moreover it has
a rank equal to 0 or 1).
• Assume that If q > 2 and that the theorem is proved up to the order q ≠ 1.
Once we reduced the �rst column (as we did in 1), one has to get the reduced row echelon form of a
matrix, denoted A1 , which is of one of these types
à í
Ü ê 1 ú ··· ú
0
.. 0
. B ..
. B
0
0
where B is a matrix with q ≠ 1 columns. According to the Induction Assumption, one can transform,
using only elementary operations on the rows, the matrix B in a reduced row echelon matrix.
By applying these elementary operations on the matrix A1 (taking into account the shifting between
line numbering of B and A1 (in the second case), we get a matrix A2 which has one of the form
below, but in which B is reduced.
In the �rst case, we are done since by adding a column full of zeros at the left of a reduced row
echelon matrix we still get a reduced row echelon matrix. In the second case, we just have to make
appear zeros on the �rst row of the matrix A2 and only for the pivot columns of matrix B. Denote
j the index of one of these columns. This column is of the form
0 10 1
– 1
B CB C
B0C B 1 C
B CB C
B .. C B 1 C
B.CB C
B CB C
B0C B
B CB 1 C
C
B CB th C.
B 1 C BΩ j rowC
B CB C
B CB
B0C B 0 C
C
B CB
B .. C B .. C
C
@.A@ . A
0 0
We therefore only have to add to the �rst row of A2 , the j th row of A2 multiplied by ≠– in order to
get the desired form for the pivot column. This does not a�ect the columns of A2 which extend the
pivot columns of B which have a di�erent index from j (but modify in general the other columns).
This achieves the proof of the theorem. ⇤
Remark 2.3.12
Let A be in Mp,q (K). In order to determine the matrices G and R := GA, one can use Gaussian
Elimination. In order to avoid having to remember the elementary operations we have done to reach
this goal, we use the fact that G is obtained by performing all these elementary operations on the
unit matrix. Thus we are going to perform the required elementary operations on the matrix A and,
in parallel, on the unit matrix Imin{p,q} . Once the matrix A has been transformed into a reduced row
echelon matrix R, Imin{p,q} will have been transformed into the matrix G.
In the next example, we show how to get the reduced row echelon form of 2 ◊ 3 matrix. In the sequel,
when it comes to get the reduced row echelon form of a given matrix, we will use extensively, the
20 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
following notation: Ri Ω≠ Ri ≠ 2Rj . It means that the new row i is the former row i, from which we
subtracted 2 times the former row j
Example 2.3.13
Ç å
2 ≠2 3
Give the row reduced echelon form of the matrix A := .
4 3 7
We start with the following position:
Ö è
1 0 Ç å Ç å
2 ≠2 3 1 0
Starting positon
4 3 7 0 1
0 1
Ö è
1 0 Ç å Ç å
2 ≠2 3 1 0
Operation S(≠2, 1, 2): i.e. R2 Ω≠ R2 ≠ 2R1
0 7 1 ≠2 1
0 1
Ö è
1 0 Ç å Ç å
1 1 ≠1 3/2 1/2 0
Operation M( 2 , 1) : i.e. R1 Ω≠ R1 /2
0 7 1 ≠2 1
0 1
Ö è
1 0 Ç å Ç å
1 ≠1 3/2 1/2 0
Operation M( 17 , 2) : i.e. R2 Ω≠ R2 /7
0 1 1/7 ≠2/7 1/7
0 1
Ö è
1 0 Ç å Ç å
1 0 23/14 3/14 1/7
Operation S(1, 2, 1) : i.e. R1 Ω≠ R1 + R2
0 1 1/7 ≠2/7 1/7
0 1
Ç å Ç å
1 0 23/14 3/14 1/7
The matrix R := is in reduced row echelon form of A and G := is
0 1 1/7 ≠2/7 1/7
the unique invertible matrix such that GA = R. Here, it is easy to verify:
Ç 3 1
å Ç å Ç 23
å
14 7 2 ≠2 3 1 0 14
G·A= · = = R.
≠ 27 1
7 4 3 7 0 1 1
7
Remark 2.3.14
Since the notation Ri Ω≠ aRi + bRj is a more instinctive way to denote the elementary operations
we are performing, when one tries to determine the reduced row echelon form, we will exclusively
use it from now on. We will therefore only write R1 Ω≠ R1 + R2 instead of Operation S(1, 2, 1) :
i.e. R1 Ω≠ R1 + R2 .
21 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
Elimination then allows not only to establish that a matrix A is invertible, by showing that R = I, but
also to compute its inverse G (since GA = R = I).
Example 2.4.1
Ö è
2 ≠2 3
Let us show that A := 4 1 3 is invertible and let us determine its inverse. Before doing so,
2 0 3
let us remind that R2 Ωæ R3 means that we interchange Row 2 and Row 3.
R2 Ω≠ R2 ≠ 4R1 Ö è Ö è
2 ≠2 3 1 0 0
Starting positon
4 1 3 0 1 0
1
2 0 3 0 0 1
1
Ö è Ö è
R1 Ω≠ R21 1 ≠1 3/2 1/2 0 0
R2 Ω≠ R2 ≠ 4R1 4 1 3 0 1 0
R3 Ω≠ R3 ≠ 7R1 2 0 3 0 0 1
Ö è Ö è
R1 Ω≠ R1 ≠ 2R2 1 ≠1 3/2 1/2 0 0
R2 Ω≠ R2 ≠ 4R1 0 5 ≠3 ≠2 1 0
R3 Ω≠ R3 ≠ 2R1 0 2 0 ≠1 0 1
Ö è Ö è
R2 Ω≠ ≠R2 /3 1 ≠1 3/2 1 0 0
R2 Ωæ R3 0 2 0 ≠1 0 1
R2 Ω≠ R2 ≠ 2R1 0 5 ≠3 ≠2 1 0
Ö è Ö è
R2 Ω≠ ≠R2 /3 1 ≠1 3/2 1 0 0
R2 Ωæ 2 R2
0 1 0 ≠1/2 0 1/2
R2 Ω≠ R2 ≠ 2R1 0 5 ≠3 ≠2 1 0
Ö è Ö è
R1 Ω≠ R1 + R2 1 0 3/2 1/2 0 1/2
R1 Ω≠ R1 ≠ R1 0 1 0 ≠1/2 0 1/2
R3 Ω≠ R3 ≠ 5R2 0 0 ≠3 1/2 1 ≠5/2
Ö è Ö è
R1 Ω≠ R1 + R3 1 0 3/2 1/2 0 1/2
R2 Ω≠ R2 ≠ 2R3 0 1 0 ≠1/2 0 1/2
R3 Ω≠ ≠ 3R3
0 0 1 ≠1/6 ≠1/3 5/6
Ö è Ö è
R1 Ω≠ R1 ≠ 32 R3 1 0 0 1/4 1/2 ≠3/4
R2 Ω≠ R2 ≠ 2R3 0 1 0 ≠1/2 0 1/2
R3 Ω≠ ≠ 3 R3
0 0 1 ≠1/6 ≠1/3 5/6
Since the reduced row echelon matrix is the unit matrix, it appears that A is invertible and that its
inverse is: Ö 1 è
1
4 2 ≠ 34
A≠1 = ≠ 12 0 1
2
.
≠ 16 ≠ 13 5
6
22 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
Ö è Ö 1 1
è
2 ≠2 3 4 2 ≠ 34
We can therefore check that 4 1 3 · ≠ 12 0 1
2
= I3 .
2 0 3 ≠ 16 ≠ 13 5
6
We end this section by the following result, which be useful in the next section.
Proposition 2.4.2
Let A be a p ◊ q matrix and F be a p ◊ p invertible matrix. Matrices A and F A have the same rank.
Proof 2.4.3
Let G be an invertible matrix such that R = GA is a reduced row echelon matrix. One can write:
GA = R ≈∆ GF ≠1 · F A = R.
Since GF ≠1 is invertible, this shows that A and F A have the same reduced row echelon matrix, and
thus the same rank. ⇤
1. only by the elementary operations described at Section 2.2, at the exception of interchang-
ing rows.
2. the pivots obtained via the Gaussian elimination algorithm are all non-zero pivots.
2
i.e. a linear system for which Matrix A is triangular
23 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
Example 2.5.4
De�ne the following matrices
Ö è Ö è Ö è
2 ≠2 3 0 2 0 0 6 0
A := 4 1 3 , C := 2 0 0 , & F := 2 1 1 .
2 0 3 0 0 2 0 1 8
Remark 2.5.5
Careful readers recognized that Matrix A, given in Example 2.5.4 is the same as the one given in
Example 2.4.1. Since, we interchanged row 2 and row 3, in the Gaussian Elimination performed in
Example 2.4.1, it might be surprising to state that A is regular. In fact, one can perform Gaussian
24 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
R2 Ω≠ R2 ≠ 4R1 Ö è Ö è
2 ≠2 3 1 0 0
Starting positon
4 1 3 0 1 0
1
2 0 3 0 0 1
1
Ö è Ö è
R1 Ω≠ R21 1 ≠1 3/2 1/2 0 0
R2 Ω≠ R2 ≠ 4R1 4 1 3 0 1 0
R3 Ω≠ R3 ≠ 7R1 2 0 3 0 0 1
Ö è Ö è
R1 Ω≠ R1 ≠ 2R2 1 ≠1 3/2 1/2 0 0
R2 Ω≠ R2 ≠ 4R1 0 5 ≠3 ≠2 1 0
R3 Ω≠ R3 ≠ 2R1 0 2 0 ≠1 0 1
Ö è Ö è
R2 Ω≠ ≠R2 /3 1 ≠1 3/2 1/2 0 0
R2 Ω≠ R2 /5 0 1 ≠3/5 ≠2/5 1/5 0
R2 Ω≠ R2 ≠ 2R1 0 2 0 ≠1 0 1
Ö è Ö è
R1 Ω≠ R1 + R2 1 0 9/10 1/10 1/5 0
R1 Ω≠ R1 ≠ R1 0 1 ≠3/5 ≠2/5 1/5 0
R3 Ω≠ R3 ≠ 2R2 0 0 6/5 ≠1/5 ≠2/5 1
Ö è Ö è
R1 Ω≠ R1 + R3 1 0 9/10 1/10 1/5 0
R2 Ω≠ R2 ≠ 2R3 0 1 ≠3/5 ≠2/5 1/5 0
5
R3 Ω≠ 6 R3 0 0 1 ≠1/6 ≠1/3 5/6
9
Ö è Ö è
R1 Ω≠ R1 ≠ 10 R3 1 0 0 1/4 1/2 ≠3/4
3
R2 Ω≠ R2 ≠ 5 R3 0 1 0 ≠1/2 0 1/2
R3 Ω≠ ≠ 3R3
0 0 1 ≠1/6 ≠1/3 5/6
The reason why we performed Gaussian elimination the way we did it, in Example 2.4.1, was to
9
simplify computations (and not get a ≠ 10 or 35 for example).
A = LU (5.1)
where L is a lower unitriangular matrix, having all 1’s on the diagonal, and U is upper triangular with
non zero diagonal entries, which are the pivots of A. The non zero o�-diagonal entries lij for i > j
appearing in L prescribe the elementary row operations that bring A into upper triangular form;
25 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
namely, one subtracts li,j times row j from row i at the appropriate step of the Gaussian Elimination
process.
Remark 2.5.7
In order to avoid any confusion between LU factorization (given at De�nition 2.5.1, and where we
do not have any information about the diagonal entries of U ) and the LU factorization described
right above (where we know that matrix U has non zero diagonal entries), we will call this latter
regular LU factorization.
In practice, to �nd the regular LU factorization of a square matrix A, one applies the regular Gaus-
sian Elimination algorithm to reduce A to its upper triangular form U . The entries of L can be �lled in
during the course of the calculation with the negatives of the multiples used in the elementary row op-
erations. If the algorithm fails to be completed, which happens whenever zero appears in any diagonal
pivot position or if we have to interchange rows to get the row echelon form of A, then the original
matrix does not have an regular LU factorization. Let’s see on an example how the LU factorization
can be performed.
Example 2.5.8
Let’s compute the LU factorization of the matrix
Ö è
1 2 1
A := 3 10 3 .
≠2 ≠8 5
At this stage we do not even know if such a factorization does exist for A. Only the Gaussian Elimi-
nation algorithm will tell us if it does. Thus, applying the Gaussian Elimination algorithm, we begin
by de�ning A1 to be A and L1 to be I3 . We hence start with:
R2 Ω≠ R2 ≠ 2R1 Ö è Ö è
1 2 1 1 0 0
R2 Ω≠ R2 ≠ 4R1
3 10 3 := A1 0 1 0 := L1
1
≠2 ≠8 5 0 0 1
1
Ö è Ö è
R1 Ω≠ R1 ≠ 2R2 1 2 1 1 0 0
R2 Ω≠ R2 ≠ 3R1 0 4 0 := A2 3 1 0 := L2
R3 Ω≠ R3 + 2R1 0 ≠4 7 ≠2 0 1
Ö è Ö è
R1 Ω≠ R1 ≠ 2R2 1 2 1 1 0 0
R2 Ω≠ R2 ≠ 2R1 0 4 0 := A3 3 1 0 := L3
R3 Ω≠ R3 + 1R2 0 0 7 ≠2 ≠1 1
Note that A3 is an upper triangular matrix while L3 is a lower triangular matrix. Thus one de�nes
U to be A3 and L to be L3 . In other words, de�ne matrices L and U by setting:
Ö è Ö è
1 0 0 1 2 1
L := 3 1 0 & U := 0 4 0 .
≠2 ≠1 1 0 0 7
Note that the diagonal entries of U are the pivots of A. Since they are all non zero here, it is clear, in
view of Theorem 2.5.6, that A is regular. Therefore the LU factorization we just gave is the regular
LU factorization of A.
26 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
Note that L has always all its diagonal elements equal to 1 while its entries lying below the main
diagonal are the negatives of the multiples we used during the elimination procedure. Moreover, it
is easy to check that:
Ö è Ö è Ö è
1 0 0 1 2 1 1 2 1
L·U = 3 1 0 · 0 4 0 = 3 10 3 = A.
≠2 ≠1 1 0 0 7 ≠2 ≠8 5
Remark 2.5.9
By trying Gaussian elimination on matrices B and C given in Example 2.5.4 we immediately see that
the algorithm requires an interchange of rows. Thus none of these two matrices are regular. As a
consequence none of them admit a regular LU factorization.
As the next subsection will show, that it is possible, if we get the rid of the condition concerning the
interchanging of rows, to extend the regular LU factorization to any invertible matrix (and in fact to a
bigger set of matrices). The price to pay will be that we will not have A = LU anymore but P A = LU ,
for a certain invertible matrix P .
P A = LU Factorization
Before starting this section, we must have in mind, what the permutation matrices are. We de�ned
them at page 14. Hence, it is easy to see that, in M3 (R), there are 3! = 6 permutations matrices that
are:
Ö è Ö è Ö è
0 1 0 0 0 1 1 0 0
P (1, 2) = 1 0 0 P (1, 3) = 0 1 0 P (2, 3) = 0 0 1
0 0 1 1 0 0 0 1 0
Ö è Ö è Ö è
0 1 0 0 0 1 1 0 0
P (2, 1, 3) = 0 0 1 P (3, 1, 2) = 1 0 0 PId = 0 1 0 .
1 0 0 0 1 0 0 0 1
Remark 2.5.11
As we will see in Theorem 2.5.16, invertible is a synonym of nonsingular, for square matrices.
A matrix which is not nonsingular is said to be singular.
We can now state the most general result, the proof of which can be found in [HJ13, Theorem 3.5.7].
27 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
Remark 2.5.13
A unit lower triangular matrix is just a lower triangular matrix the diagonal entries of which are all
equal to 1.
Remark 2.5.15
The main information added when one goes from P A = LU Factorization to nonsingular P A =
LU Factorization is that, in this latter case all the diagonal entries of U are non zero.
1. A is nonsingular.
2. A is invertible.
Proof 2.5.17
1 =∆ 3: by very de�nition of non singular matrices.
3 =∆ 1: is a straight consequence of Gaussian elimination.
3 =∆ 2: Since A has p non zero pivots, the Matrix R, provided by the Gaussian elimination of A,
must equals Ip . Moreover, since the equality GA = R holds, this proves that A is invertible and that
A≠1 = G.
2 =∆ 3: Let us prove this result by using the contrapositive. In other words let us prove that
non 3 =∆ non 2.
Denote G and R the matrices provided by the Gaussian elimination of A. We hence have:
GA = R (5.2)
If A does not have p nonzero pivots, then R can not be Ip . In particular R is not invertible.
Let’s assume that A is invertible. In this case GA itself is invertible, according to Proposition 1.6.5.
Thus, one can write
(5.2) ≈∆ (GA)≠1 · GA = (GA)≠1 R ≈∆ Ip = (GA)≠1 R.
In other words, R is invertible. This contradicts the fact that R is not invertible. Thus our assumption
that A is invretible does not hold.
28 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
Example 2.5.18
Let us give a permuted LU factorization of the matrix
à í
1 2 ≠3 4
≠5 ≠10 9 ≠36
A := .
≠2 ≠6 5 ≠5
3 0 ≠9 27
We hence start with:
à í à í à í
1 2 ≠3 4 1 0 0 0 1 0 0 0
≠5 ≠10 9 ≠36 0 1 0 0 0 1 0 0
:= A1 := L1 := P1
≠2 ≠6 5 ≠5 0 0 1 0 0 0 1 0
3 0 ≠9 27 0 0 0 1 0 0 0 1
To begin the procedure, we eliminate the entries below the �rst pivot.
à í à í à í
R2 Ω≠R2 +5R1 1 2 ≠3 4 1 0 0 0 1 0 0 0
R2 Ω≠R2 +5R1 0 0 ≠6 ≠16 ≠5 1 0 0 0 1 0 0
:=A2 :=L2 :=P2 .
R3 Ω≠R3 +2R1 0 ≠2 ≠1 3 ≠2 0 1 0 0 0 1 0
R4 Ω≠R4 ≠3R1 0 ≠6 0 15 3 0 0 1 0 0 0 1
Since the (2, 2) entry of A2 is zero, we interchange rows 2 and 3 which leads us to:
à í à í à í
1 2 ≠3 4 1 0 0 0 1 0 0 0
R2 Ω≠R2 ≠2R1
0 ≠2 ≠1 3 ≠2 1 0 0 0 0 1 0
R2 ΩæR3 :=A3 :=L3 :=P3 .
0 0 ≠6 ≠16 ≠5 0 1 0 0 1 0 0
R4 Ω≠R2 +1R1
0 ≠6 0 15 3 0 0 1 0 0 0 1
We interchanged the same two rows of P , while in L we only interchanged the already computed
entries in tis second and third rows that lie in its �rst column below the diagonal. We then eliminate
the nonzero entry lying below the (2, 2) pivot, leading to:
à í à í à í
R2 Ω≠R2 ≠2R1 1 2 ≠3 4 1 0 0 0 1 0 0 0
R2 Ω≠R3 0 ≠2 ≠1 3 ≠2 1 0 0 0 0 1 0
:=A4 :=L4 :=P4 .
R3 Ω≠R2 0 0 ≠6 ≠16 ≠5 0 1 0 0 1 0 0
R4 Ω≠R4 ≠3R2 0 0 3 6 3 3 0 1 0 0 0 1
29 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
We thus de�ne
U := A6 , L := L6 , & P := P6 .
U having all non zero pivots, it is, as well as A, non singular. Besides, one can verify that:
à í à í
1 0 0 0 1 2 ≠3 4
0 0 1 0 ≠5 ≠10 9 ≠36
PA = ·
0 0 0 1 ≠2 ≠6 5 ≠5
0 1 0 0 3 0 ≠9 27
à í à í à í
1 2 ≠3 4 1 0 0 0 1 2 ≠3 4
≠2 ≠6 5 ≠5 ≠2 1 0 0 0 ≠2 ≠1 3
= = ·
3 0 ≠9 27 3 3 1 0 0 0 3 6
≠5 ≠10 9 ≠36 ≠5 0 ≠2 1 0 0 0 ≠4
= LU. (5.3)
Remark 2.5.19
1. In order not to overload notations one usually does not specify the nature of the LU factor-
ization (i.e. regular or nonsingular nor if it is or not a permuted one) since it is clear from the
context.
3. Note that when one deals with a non invertible matrix A, a “kind” of permuted LU factorization
exists but there will be some zeros on the diagonal of U (as shows Example 2.6.15 below).
30 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
Readers interested in more details about LU factorization (in particular, when A is singular) should
refer to [HJ13, p.160-163], and in particular to [HJ13, Theorem 3.5.7 p.163]. The �gure below gives a
summary of how the regular, nonsingular and singular matrices are linked to each others.
Proposition 2.6.1
A system of p linear equations with q unknowns has either 0, or 1 or in�nitely many solutions. If
q > p, it has either 0, or in�nitely many solutions.
Proof 2.6.2
The linear system we are studying has p equations and q unknowns and can be written, under the
matrix form AX = B. According to Gauss theorem, there exists a unique invertible matrix G such
that R = GA is a reduced row echelon matrix. We therefore just have to solve the system RX = GB,
which is equivalent to the initial system. Denote r the rank of the reduced row echelon matrix R.
Since the p ≠ r last rows of matrix R equal zero, the coe�cients of the vector RX all equal 0, for
r < i 6 p. Two cases are therefore possible.
- Coe�cients (GB)i of he column vector GB are not all equal to 0, for r < i 6 p. There is an
equation of the system which has the form
0x1 + · · · + 0xq = –
31 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
- Coe�cients (GB)i of he column vector GB are all equal to 0, for r < i 6 p. Thus the p ≠ r
last rows of the system have the form
0x1 + · · · + 0xq = 0.
By cancelling them out, we get a system (of r equations with q unknowns), which is clearly
equivalent to the initial system. Denote j1 , . . . , jr the columns indices of R that contain the
pivots of R. By sending the unknowns that correspond to the columns of R that do not contain
the pivots of R, in the second member of the equation, the ith equation of the system becomes:
ÿ
xji = (GB)i ≠ Rij xj .
16j6q
j”={j1 ,...,jr }
The system therefore gives the r unknowns xj1 , . . . , xjr when one knows the q ≠ r others.
• If r = q, the unknowns are all determined and the system has only one solution, that is
((GB)1 , . . . , (GB)q ).
• If r < q, one can arbitrary choose q≠r unknowns and the system has in�nitely many solutions.
• Particular case p < q: Since r 6 p, one can not have r = q and the system can therefore not
have only one solution.
Corollary 2.6.4
An homogeneous linear system (i.e. with a second member equal to (0)) with p equations and q
unknowns has a non zero solution if and only if the rank of its matrix is smaller than q. This is in
particular the case if p < q.
Proof 2.6.5
The system AX = 0Rq has at least the null solution, i.e. 0Rq . It will have a non zero solution if and
only if it has in�nitely many of them. According to the proof of Proposition 2.6.1, it happens if and
only if r < q. ⇤
The results stated in the previous section remain valid. However, one can be more accurate when it
comes to square linear systems.
From now on, de�ne Let A be a matrix of Mp (K) and B be a column matrix with p rows. Denote
32 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
à í
x1
x1
X := .. and de�ne the following system:
.
xp
(S) AX = B.
The next result gives the methodology to solve such a linear system if A is invertible, using the inverse
of A.
Theorem 2.6.6 (Characterization of Invertible Matrices)
Considering the System (S) described above, the following conditions are equivalent:
1. A is an invertible matrix.
2. The rank of A equals p (in other words the reduced matrix, associated to A has no null row).
4. For all column matrix B, the system AX = B has at least one solution.
Moreover, when one of these conditions holds, then the reduced row echelon matrix, associated to
A, is Ip and the unique solution of the system AX = B is ((A≠1 B)1 , (A≠1 B)2 , · · · , (A≠1 B)p ).
Proof 2.6.7
1 =∆ 5 is a straight consequence of Proposition 2.4.2.
Proof 2.5.17 already established the equivalence 1 ≈∆ 2.
5 =∆ 4 and 5 =∆ 3 are obvious.
It remains to prove 4 =∆ 2 and 3 =∆ 2. We will use the contrapositive and show that non 2 implies
both non 3 and non 4.
Denote G and R the matrices associated to A by Gauss theorem. We hence have GA = R, and R is
a reduced row echelon matrix.
non 2 =∆ non 4:
non 2 means that the rank of A is not equal to p. It is therefore smaller than p. Hence the coe�cient
on the last row of R are all equal to 0. Let B be a column matrix such that the coe�cient, on its last
row, equals 1. System AX = G≠1 B is equivalent to the system RX = B. Since the last equation of
this latter system is:
0x1 + · · · + 0xp = 1,
it has no solution. Hence Condition 4 is false. This proves non 2 =∆ non 4.
non 2 =∆ non 3:
Homogeneous system AX = (0) is equivalent to the system RX = (0). In this latter, one can
suppress the last equation, which is of the form:
0x1 + · · · + 0xp = 0.
We get an equivalent system, which has p≠1 equations and p unknowns. Since it has more unknowns
than equations, Proposition 2.6.1 tells us that it can not have a unique solution. In other words 3 is
not true. This proves non 2 =∆ non 3 and achieves the proof. ⇤
33 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
Corollary 2.6.8
Let A be a matrix in Mp (K). The following conditions are equivalent:
1. A is an invertible matrix.
Proof 2.6.9
1 =∆ 2 and 1 =∆ 3 are obvious.
2 =∆ 1:
Assume that 2 holds. Let X be a column matrix such that AX = 0Rp . By multiplying, from the left
hand side by G, we get: GAX = 0Rp i.e. X = 0Rp . This shows that 0Rp is the only solution of
the system. Condition 3 of Theorem 2.6.6 therefore holds, and thus, establishes that Condition 1 of
Theorem 2.6.6 holds as well.
3 =∆ 1:
Assume that 3 holds. Let B be a column. De�ne X := DB. By multiplying, from the left hand side
by A, we get: AX = ADB = B. Hence Condition 4 of Theorem 2.6.6 holds, and thus, establishes
that Condition 1 of Theorem 2.6.6 holds as well. ⇤
The next result will give a justi�cation to the methodology for solving linear systems, when A is
invertible, using the P A = LU factorization.
Proof 2.6.11
The proof 2.6.7 shows that there is a unique solution to such a system. The fact that one can use
permuted LU -factorization to solve the system becomes clear once one reads the following method-
ology. ⇤
P AX = P B =: B Õ . (6.4)
Note that B Õ has been obtained by permuting the entries of B in the same fashion as the rows of A.
Moreover (6.4) can be rewritten under the form:
LU X = B Õ . (6.5)
LC = B Õ , (6.6)
34 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
where C is the unknown vector in (6.6), and is solved by Forward Substitution. The second triangular
system to solve, by Backward Substitution this time, is
UX = C (6.7)
in the unknown X. Let’s apply this methodology in the following example.
Example 2.6.12
Let us solve the following linear system:
(T) AX = B,
where
à í à í
1 2 ≠3 4 1
≠5 ≠10 9 ≠36 ≠1
A := and B := .
≠2 ≠6 5 ≠5 3
3 0 ≠9 27 0
According to the conclusion of Theorem 2.6.6, we therefore know that X = A≠1 B. Thus
0 1
(A≠1 B)1 à í à í à í
231 15 51 53
B
B
C
C ≠ 4 4 ≠ 2 6 1 ≠138
B (A≠1 B)2 C 69
≠ 58 15
≠ 17 ≠1 41
B C 8 4 12 2
X=B C= 19 1 4
· = (6.8)
B (A≠1 B) C
B 3 C
≠ 2 2 ≠4 3 3 ≠22
@ A 13
≠4 1 3
≠ 12 0 8
4 2
(A B)4
≠1
De�ne Sol(T) the set of solutions of System (T). We just proved that:
¶Ä ä©
Sol(T) = 41 ,
t
≠138 2 ≠22 8
The P A = LU factorization of A has been computed at (5.3) (on page 30) and is therefore not recalled
here. According to the methodology given right after Theorem 2.6.10, de�ne
à í à í
1 0 0 0 1
0 0 1 0 3
B Õ := ·B = .
0 0 0 1 0
0 1 0 0 ≠1
35 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
0 1
c1
B C
B C
B c2 C
B C
Denote C := B C. We know that we �rst need to solve, by forward substitution, the system
B c C
B 3 C
@ A
c4
LC = B Õ , (6.9)
U X = C, (6.11)
Remark 2.6.13
When it comes to solving a square linear systems AX = B for several di�erent matrices B, the per-
muted LU factorization can be repeatedly applied to solve the equation multiple times for di�erent
B. In this case it is faster (and more convenient) to do a permuted LU decomposition of the matrix
A once and then solve the triangular matrices system for the di�erent B, rather than using Gaussian
elimination. Indeed, inverse matrices are very convenient in analytical manipulations, because they
allow you to move matrices from one side to the other of equations easily. However, inverse matrices
are almost never computed in “serious” numerical calculations. Whenever you see A≠1 B, when you
36 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
go to implement it on a computer you should read A≠1 B as “solve AX = B by some method”: e.g.
solve it by �rst computing the LU factorization of A and then using it to solve AX = B.
One reason that you don’t usually compute inverse matrices is that it is wasteful: once you have
P A = LU , you can solve AX = B directly without bothering �nding A≠1 , and computing A≠1
requires much more work if you only have to solve a few right-hand sides. Another reason is that for
many special matrices, there are ways to solve AX = B much more quickly than you can �nd A≠1 .
For example, many large matrices in practice are sparse (mostly zero), and often for sparse matrices
you can arrange for L and U to be sparse too. Sparse matrices are much more e�cient to work with
than general “dense” matrices because you don’t have to multiply (or even store) the zeros. Even if
A is sparse, however, A≠1 is usually non-sparse, so you lose the special e�ciency of sparsity if you
compute the inverse matrix.
The matrices L and U could be thought to have “encoded” the Gaussian elimination process.
Moreover, The cost of solving a system of linear equations is approximately 23 p3 �oating-point op-
erations if the matrix A belngs to Mp (R). This makes it twice as fast as algorithms based on QR
decomposition, which costs about 43 p3 �oating-point operations when Householder re�ections are
used. For this reason, LU decomposition is usually preferred. The readers interested in these nu-
merical details may refer to [Str16, p.101-102].
The next theorem summarizes the ways of solving a square linear system. The proof is obvious, in
view of Theorem 2.6.6 and Theorem 2.6.10.
(S) AX = B,
where we denoted X := t (x1 , x2 , · · · , xp ) the unknown vector. Depending on the rank of the matrix
A, there are di�erent possible issues when one wants to solve (S).
• One can solve (S) by invoquing Point 5. of Theorem 2.6.6. The unique solution of the system
AX = B is ((A≠1 B)1 , (A≠1 B)2 , · · · , (A≠1 B)p ). One just have to compute the inverse of A.
• One can also solve (S) using the P A = LU method (see Theorem 2.6.10 and the methodology
given right below it).
Let us �nish this chapter with an example of linear system, the solution of which is not unique.
37 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
Example 2.6.15
We want to solve the following linear system.
8
< x≠z =1
>
(S ) : 2y + 3z = 0 .
>
:
≠2x + 2z = ≠2
To this end, de�ne
á ë á ë
Ö è 1 x
1 0 ≠1
A := 0 2 3 B := 0 , & X := y .
≠2 0 2
≠2 z
It is clear that (S ) is equivalent to AX = B. Since the last row of A is proportional to the �rst one,
it is clear that A is not invertible (since its rank is smaller than 3).
1. Using the P A = LU factorization.
One could either use Gaussian elimination or the permuted LU factorization of A to solve (S ). Let
use �rst the permuted LU factorization of A. We hence start with:
Ö è Ö è Ö è
1 0 ≠1 1 0 0 1 0 0
0 2 3 := A1 0 1 0 := L1 0 1 0 := P1
≠2 0 2 0 0 1 0 0 1
The second and last step is therefore:
Ö è Ö è Ö è
1 0 ≠1 1 0 0 1 0 0
0 2 3 := A2 0 1 0 := L2 0 1 0 := P2
0 0 0 ≠2 0 1 0 0 1
We thus de�ne
U := A2 , L := L2 , & P := P2 .
We then have to solve, �rst:
á ë
Ö èÖ è 1
1 0 0 c1
LC = P B i.e. 0 1 0 c2 = 0
≠2 0 1 c3
≠2
á ë
1
The solution of such a system is, obviously, C = 0 . We then have to solve:
0
á ë
Ö èÖ è 1
1 0 ≠1 x1
UX = C i.e. 0 2 3 x2 = 0
0 0 0 x3
0
38 S����� 2023
L����� A������ Lecture Notes C������ 2: Linear Systems
We see easily see that (S Õ ) (and thus (S )) has in�nitely many solutions. More precisely, if S ol(S )
denotes the set of solutions of (S ), one can write:
⁄
S ol(S ) = {(1 + –, ≠3–
2 , –), – œ R}.
We just need to get the reduced row echelon form of A. A few computations provides us with:
Ö è Ö è
1 0 ≠1 1 0 0
R := 0 2 3 & G := 0 1/2 0
0 0 0 ≠2 0 1
Since we know, by de�nition of reduced row echelon form, that GA = R, one can write:
(S ) ≈∆ AX = B ≈∆ GAX = GB ≈∆ RX = GB
á ë á ë
Ö è x Ö è 1
1 0 ≠1 1 0 0
≈∆ 0 2 3 y = 0 1/2 0 0
0 0 0 ≠1 0 1
z ≠2
Such a system also reads:
®
x≠z =1
(S ÕÕ )
2y + 3z = 0
It is clear that (S ÕÕ ) is nothing but (S Õ ), with (x, y, z) instead of (x1 , x2 , x3 ). The set of solution
of (S ÕÕ ) is therefore S ol(S ) = {(1 + –, ≠3– 2 , –), – œ R}. We recovered the result from the P LU
factorization.
39 S����� 2023
Part II
40
C������ 3
Vector Spaces
K◊E æ E
(–, x) ‘æ – · x
(E, +, ·) is said to be a Vector Space (or a Linear Space) on K or a K-Vector Space (or a K-Linear
Space) if:
5. – · (x + y) = – · x + – · y
6. (– + —) · x = – · x + — · x
7. 1 · x = x
8. – · (— · x) = (–—) · x.
41
L����� A������ Lecture Notes C������ 3: Vector Spaces
ÇÇ
Ç åå Ç å å Ç å Ç å
y1 x1 –x1 + –y1 –x1 –y1
5. – · (x + y) = – · + = = + = – · x + – · y,
y2 x2 –x2 + –y2 –x2 –y2
Ç å Ç å Ç å Ç å
(– + —) · x1 –x1 + —x1 –x1 —x1
6. (– + —) · x = = = + = – · x + — · x,
(– + —) · x2 –x2 + —x2 –x2 —x2
Ç å Ç å
x1 1 · x1
7. 1 · x = 1 · = = x,
x2 1 · x2
åå Ç Ç Ç å Ç å Ç å Ç å
x1 — · x1 – · —x1 –—x1 (–—) · x1
8. – · (— · x) = – · — · =–· = = =
x2 — · x2 – · —x2 –—x2 (–—) · x2
Ç å
x1
= (–—) · = (–—) · x.
x2
42 S����� 2023
L����� A������ Lecture Notes C������ 3: Vector Spaces
Starting from the axioms introduced in 3.1.1, we easily deduce the following properties:
Proposition 3.1.1. 1. ’x œ E, 0 · x = 0E ,
2. ’⁄ œ K, ⁄ · 0E = 0E ,
3. ’x œ E, (≠1) · x = ≠x,
Proof 3.1.3
To be written. ⇤
Remark 3.1.4
1. When there is no ambiguity on K one says Vector Space instead of K-Vector Space.
2. We will often write 0 instead of 0E . The same symbol will hence denotes, from one hand the
neutral element for addition in E (and in this case 0 belongs to E) and, in an other hand, the
neutral element for addition in K (and in this case 0 belongs to K).
3. Let X be a set and let F(X, K) be the set of K-valued functions, de�ned on X. Then
(F(X, K), +, ·) is a K-Vector Space if the two laws + and · are de�ned by setting:
f +g : X æ K –·f : X æ K
t ‘æ f (t) + g(t) t ‘æ –f (t)
4. ({u := (un )nœN , un œ K}, +, ·) is a K-Vector Space The set of all sequences of real or
complex numbers, endowed with the addition between sequences and with the multiplication
of a sequence by a scalar is Vector Space on K.
43 S����� 2023
L����� A������ Lecture Notes C������ 3: Vector Spaces
6. (Mp,q (K), +, ·) is a K-Vector Space. The set of matrices with p rows and q columns, endowed
with the addition between two matrices and with the multiplication by a scalar is Vector Space
on K.
Remark 3.1.6
When there is no risk of confusion we say that E is a vector space instead of (E, +, ·).
1. 0E œ F ,
2. ’(x, y) œ F 2 , x + y œ F ,
3. ’⁄ œ K, ’x œ F, ⁄ · x œ F ,
Remark 3.2.2
It is clear that a Vector subspace is a vector space (for the laws + and · de�ned on E).
Example 3.2.3
1. R and iR are vector subspaces of the R-vector space C.
3. When F is a sub vector space of E then (F(F, K), +, ·) is vector subspace of (F(E, K), +, ·).
4. General theorems on continuity show that the set of continuous R-valued functions de�ned
on [0, 1] (denoted C 0 ([0, 1], R)) is a vector subspace of (F([0, 1], R), +, ·)
qn
5. The set F := {x := (x1 , x2 , · · · , xn ) œ Rn , i=1 xi = 0} is a vector subspace of Rn .
qn
6. The set F := {x := (x1 , x2 , · · · , xn ) œ Rn , i=1 xi = 1} is not a vector subspace of Rn
since it does not contain 0Rn .
44 S����� 2023
L����� A������ Lecture Notes C������ 3: Vector Spaces
Proposition 3.3.1
An intersection of Vector subspaces of E is a vector subspace of E.
Proof 3.3.2
To be written. ⇤
Proposition 3.3.3
The union of two vectors subspaces of E that are not included in each other is not a vector subspace
of E.
Proof 3.3.4
To be �lled To be written. [CPY96, p.48] ⇤
Remark 3.3.5
One can summarize the previous proposition by writing that an union of Vector subspaces is not, in
general, a vector subspace. For example the union of two lines in R2 is not a vector subspace of R2 .
Remark 3.3.7
One might want to use sometimes the notation
ÿ
⁄a a.
aœA
This is not ambiguous if A is a �nite subset of E but this suggests, if A is an in�nite subset of E, that
only �nitely many scalars ⁄a are di�erent from 0. Hence the given sum is correctly de�ned.
45 S����� 2023
L����� A������ Lecture Notes C������ 3: Vector Spaces
If A := (ai )iœI is a family of vectors of E, a linear combination of vectors of A is a sum of the form
q
iœI ⁄i ai in which all the ⁄i equal 0, except �nitely many of them. When we index a subset A by its
own elements (i.e. I = A), we recover the notation introduced at remark ??.
Example 3.3.10
When A = {a} has a unique element, Span(A) is the set Ka = {⁄a, ⁄ œ K} of “multiples” of the
vector a. A set E := Ka, where a ”= 0 is called a line of E.
The next proposition shows, in particular, that the lines of E are vector subspaces. This can be checked
directly.
Proposition 3.3.11
Let A be a subset of E (or a family of vectors of E). The set Span(A) is a vector subspace of E.
Moreover, it is the smallest vector subspace which contains A. It is also the intersection of all vector
subspaces of E which contain A. It is called the vector subspace spanned by A and we denote it
SpanK (A) or SpanK A (or only Span(A) or Span A when there is no risk of confusion).
Proof 3.3.12
To be written [CPY96, p.49]. ⇤
Finally, one easily deduces, from Propositions 3.3.1 and 3.3.11 that SpanK (A) is the intersection of all
vector subspaces of E which contain A.
Example 3.3.13
1. If A = ÿ, then Span(A) = {0E }.
• SpanR ({1}) = R,
• SpanR ({i}) = iR,
• SpanR ({1, i}) = C.
5. (1, X, X 2 , · · · , X n ) is a spanning family of Kn [X] since any polynomial with degree smaller
than n can be written under the form:
n
ÿ
ak X k with, forall k œ J0, nK, ak œ K.
k=0
We will see in Proposition 3.4.1 an easier way to prove Statement 5. of the previous example.
46 S����� 2023
L����� A������ Lecture Notes C������ 3: Vector Spaces
Proof 3.3.15
To be �lled ⇤
2. Let F, G and H be three sub-vector spaces of a vector space E. The smallest vector subspace
which contains F, G and H is:
F + G + H := {x + y + z, (x, y, z) œ F ◊ G ◊ H}.
Proposition 3.3.18
Let F and G be two vector subspaces of a vector space E that are in direct sum. For all vector x of
m
F G there exists a unique vector y in F and a unique vector z in G such that x = y + z.
Proof 3.3.19
To be written! [CPY96, p.50]. ⇤
47 S����� 2023
L����� A������ Lecture Notes C������ 3: Vector Spaces
Remark 3.3.21
When F and G are vector subspaces of a vector space E, to prove that F fl G = {0}, it is su�cient
to prove that x œ F fl G =∆ x = 0.
• The two vector subspaces F := Span{(1, 0)} and G := Span{(0, 1)} are supplementary
vector subspaces of K2 . Indeed, any element (x, y) in K2 can be written uniquely under
the form –(1, 0) + —(0, 1), with – = x and — = y.
• The two vector subspaces F := Span{(1, 0)} and H := Span{(1, 1)} are supplementary
vector subspaces of K2 . Indeed,
• ’(x, y) œ K2 , (x, y) = (x ≠ y) · (1, 0) + y · (1, 1) = (x ≠ y, 0) + (y, y),
• If (x, 0) = (y, y), then y = 0 and thus x = 0.
2. In K3 , the vector subspaces K2 ◊{0} and Span{(0, 0, 1)} are supplementary vector subspaces.
3. In E := F(R, R), one can show that the set E of even functions and the set O are supple-
m
mentary vector subspaces (thus E O = F(R, R)).
are two supplementary vector subspaces. Use Euclidean division to prove it!
One can generalize De�nition ?? to the case of any �nite number of vector subspaces of E.
E1 + E2 + · · · + En = Span(E1 fi E2 fi · · · fi En ).
48 S����� 2023
L����� A������ Lecture Notes C������ 3: Vector Spaces
Proposition 3.3.24
E1 + E2 + · · · + En is the set of all vectors of E which can be written under the form of a sum
x1 + x2 + · · · xn , where xi belongs to Ei , for all i in J1, nK.
Proof 3.3.25
To be written! [CPY96, p.50]. ⇤
Example 3.3.26
A family of vectors (x1 , x2 , · · · , xn ) spans E if and only if E = Kx1 + Kx2 + · · · + Kxn .
E = Span {a1 , a2 , · · ·, an }
i.e. if every element of E can be written as a linear combination of elements of the family
(a1 , a2 , · · ·, an ). In other words:
n
ÿ
’x œ E, ÷(⁄1 , ⁄2 , · · · , ⁄n ) œ Kn such that x = ⁄i ai .
i=1
Proof 3.4.2
Obvious in view of De�nitions 3.3.6 and 3.3.9. ⇤
49 S����� 2023
L����� A������ Lecture Notes C������ 3: Vector Spaces
Remark 3.4.3
1. If a family of vectors spans E, any family of vectors obtained by permuting its elements also
spans E. It does not depend on the order of the vectors which belong to the spanning family.
3. When the family of vectors one considers in in�nite one can not use Proposition 3.4.1 anymore.
One has to come back to De�nition 3.3.9.
Example 3.4.4
A family of vectors (x1 , x2 , · · · , xn ) spans E if and only if E = Kx1 + Kx2 + · · · + Kxn .
Proposition 3.4.5
Let G be a family of vectors which spans a vector space E. A family X of elements of E spans E if,
and only if, every element of G is a linear combination of elements of X.
Proposition 3.4.6
Let F and G be two vector subspaces of E. A vector x belongs to Span(F fi G) if and only if there
exists a vector y of F and a vector z of G such that x = y + z.
Proof 3.4.7
To be written [CPY96, p.50]. ⇤
1. We say that (xi )iœJ1,nK is a linearly independent family of vectors of E if, for all family
(⁄i )iœJ1,nK in Kn , the following implication holds:
n
ÿ
⁄i xi = 0 =∆ ⁄1 = ⁄2 = · · · ⁄n = 0.
i=1
2. In case the previous implication does not hold, i.e. if one can �nd a family (⁄i )iœJ1,nK in Kn such
q
that not all the ⁄i equal 0, and such that ni=1 ⁄i xi = 0, one says that the family (xi )iœJ1,nK is
linearly dependent or not linearly independent.
Remark 3.4.9
If a family of vectors is linearly independent, any subset of vectors of this family obtained by per-
muting its elements is also linearly independent. The linear independence of a family of vectors does
not depend on the order of its elements.
50 S����� 2023
L����� A������ Lecture Notes C������ 3: Vector Spaces
1. A family with only one element, denoted x, is linearly independent if and only if x ”= 0.
Indeed,
• If x = 0, then 1 · x = 0 and 1 ”= 0, and thus the family (x) is not linearly independent.
• If x ”= 0, then ⁄ · x = 0 =∆ ⁄ = 0.
’(a, b) œ R2 , a + ib = 0 =∆ a = b = 0.
6. The empty family, i.e. with no vector in it, is linearly independent since there is no linear
combination with all non zero coe�cient which equals 0.
Proposition 3.4.11
1. Every subset of vectors of a linearly independent family of vectors is linearly independent.
2. Every set of vectors containing a non linearly independent family of vectors is not linearly
independent.
Proof 3.4.12
To be written [CPY96, p.50]. ⇤
Remark 3.4.13
1. A linearly independent family of vectors can not contain the null vector.
51 S����� 2023
L����� A������ Lecture Notes C������ 3: Vector Spaces
Proposition 3.4.14
Let (xi )iœJ1,nK be a linearly independent family of vectors of E. The family of vectors
(x1 , x2 , · · · , xn , x) is linearly dependent if and only if x is a linear combination of x1 , x2 , · · · , xn .
Proof 3.4.15
To be written. ⇤
Theorem 3.4.16
Let (xi )iœJ1,nK be a family of linearly independent vectors of E and let (⁄1 , · · · , ⁄n ) and (µ1 , · · · , µn )
be two family of scalars, we have the following implication:
n
ÿ n
ÿ
⁄ i xi = µi xi =∆ ’i œ {1, 2, · · · , n}, ⁄i = µi .
i=1 i=1
Proof 3.4.17
To be written. ⇤
We also say that u is a morphism between vector spaces E and F . We say that u is an:
• endomorphism if E = F ,
• isomorphism if u is bijective,
Notations The set of linear maps from E to F is denoted L(E, F ) or LK (E, F ) if one wants to
precise the body K. Moreover, L(E) := L(E, E) denotes the set of endomorphisms of E.
52 S����� 2023
L����� A������ Lecture Notes C������ 3: Vector Spaces
Remark 3.5.2
If u : E æ F is a linear map then
2.
’x œ E, u(≠x) = ≠u(x).
3. We immediately can prove (by induction on n in N) that, for all (x1 , x2 , · · · , xn ) in E n and
(⁄1 , ⁄2 , · · · , ⁄n ) in Rn :
n
ÿ n
ÿ
u( ⁄ i xi ) = ⁄i u(xi ).
i=1 i=1
4. The di�erentiation is a linear map from C 1 (R) in C 0 (R). This means that:
 : C 1 (R) æ C 0 (R)
f ‘æ f Õ
: C 2 (R) æ C 0 (R)
f ‘æ af ÕÕ + bf Õ + cf
is a linear map.
6. If F and G are supplementary vector subspaces of E, one can check that the map, from E to
F , “projection on F , parallel to G” is linear.
2. On the segment [a, b], the Riemann Integral is a linear form on C 0 (I). Indeed, it is easy to
53 S����� 2023
L����� A������ Lecture Notes C������ 3: Vector Spaces
verify that:
: C 0 (I) æ R is linear.
sb
f ‘æ a f (t) dt
is linear.
Proposition 3.5.6
Let f and g be two endomorphisms of E and – and — be two scalars. Then the map –f + —g is linear.
Proposition 3.5.7
1. If f œ L(E, F ) and g œ L(F, G) then g ¶ f belongs to L(E, G).
Proof 3.5.8
To be written! [CPY96, Proposition II.17 p.52]. ⇤
are linear.
3. (L(E), +, ¶) is non commutativea in general. In other words, for u and v in L(E), u¶v ”= v¶u
in general .
a
For those who have notions of algebra, one says that (L(E), +, ¶) is a non commutative ring.
54 S����� 2023
L����� A������ Lecture Notes C������ 3: Vector Spaces
Proposition 3.5.11
(GL(E), ¶) is a group. We call it the Linear Group of E.
u(E Õ ) := {u(x), x œ E Õ },
u≠1 (F Õ ) := {x œ E, u(x) œ F Õ }.
Proposition 3.5.12
Let u : E æ F be a linear map.
Remark 3.5.14
Note that one can also denote N (u) the kernel of u.
Proposition 3.5.15
Let u : E æ F be a linear map. Then the set Ker(u) is a vector subspaces of E and the set Im(u) is
a vector subspaces of F .
55 S����� 2023
L����� A������ Lecture Notes C������ 3: Vector Spaces
Proof 3.5.16
To be �lled ⇤
between .tex
Theorem 3.5.17
Proof 3.5.18
To be �lled ⇤
Remark 3.5.19
1. When one knows that a map is linear, in order to show that it is injective one usually uses
Property 2. or its translation 3.
3.6 Basis
The following example gives basis of some classical vector spaces. The linear independence of the
family of vectors presented below has been established in Example ??, on page ??.
2. The family constituted with vectors e1 := (1, 0, 0), e2 := (0, 1, 0) and e3 := (0, 0, 1) forms a
basis of K3 . Note that B := (e1 , e2 , e3 ) is called the standard basis of K3 . Moreover, any
vector x := (x1 , x2 , x3 ) of K3 can be written as x = x1 e1 + x2 e2 + x3 e3 . Thus (e1 , e2 , e3 )
3. More generally, we call standard basis of Kn the family of vectors (e1 , e2 , · · · , en ) where,
for every k in J1, nK, ek := (0, 0, · · · , 0, 1, 0, · · · , 0) i.e. all entries are 0 except for the k th
which equals 1.
qn
4. The family (1, X, X 2 , · · · , X n ) is a basis of Kn [X] since every single polynomial k=0 ⁄k Xk
equals 0 if and only if all ⁄k equal 0. it is called the standard basis of Kn [X].
56 S����� 2023
L����� A������ Lecture Notes C������ 3: Vector Spaces
6. Let p and q be two positive integers. Denote Ei,j the p ◊ q matrix the coe�cients of which
are all zeros except for the one on the ith row and j th column which equals 1. Let A :=
(ai,j )(i,j)œJ1,pK◊J1,qK be a matrix in Mp,q (K). We clearly have the equality:
p ÿ
ÿ q
A= aij Ei,j .
i=1 i=1
One then deduces that the family (Ei,j )(i,j)œJ1,pK◊J1,qK is a basis of Mp,q (K).
Remark 3.6.3
The two following formula might be of interest for some exercises we will perform later.
®
2 (0) if j ”= k
’(i, j) œ J1, pK , t
Ei,j = Ej,i & Ei,j · Ek,l =
Ei,l if j = k
The family (⁄i )iœJ1,nK is called family of components or coordinates of x in the basis B := {ei , i œ
J1, nK}.
Proof 3.6.5
To be �lled ⇤
Remark 3.6.6
For any given family family of vectors (x1 , x2 , · · · , xn ) of E, one can easily prove that the map:
Kn æ E
qn
(⁄i )iœJ1,nK ‘æ i=1 ⁄i xi
is linear. Moreover,
• it is injective (or one to one) if, the family (xi )iœJ1,nK is linearly independent.
57 S����� 2023
L����� A������ Lecture Notes C������ 3: Vector Spaces
Proposition 3.6.7
Let u : E æ F be a linear map and let A := (ai )iœI be a family of vectors of E.
1. If u is injective then the family A is linearly independent in E if and only if (u(ai ))iœI is
linearly independent in F .
2. If u is surjective then the family A spans E if and only if (u(ai ))iœI spans F .
3. If u is bijective then the family A is a basis of E if and only if (u(ai ))iœI is a basis of F .
Proposition 3.6.8
To be written!
’i œ I, u(ai ) = bi .
Moreover,
• v is vector of F ,
58 S����� 2023
L����� A������ Lecture Notes C������ 3: Vector Spaces
where the functions a, b and c are all given and continuous over an interval I, and where the
unknown is the map y, is a linear system. Indeed, one just have to de�ne u in the following
manner:
u : C 1 (I) æ C 0 (I)
y ‘æ a.y Õ + b.y
Equation (7.3) can then be written as:
u(y) = c.
u(x) = 0F . (E0 )
Equation (E0 ) is called homogeneous equation (or equation without second member) associated to
(E). We have the following result.
2. If S(E) ”= ÿ, and if x0 is an element of S(E) , then the set of all solutions of (E) is:
In other words, the set of all solutions of (E0 ), if it is not empty, is an a�nea subspace of E.
a
This geometric notion will be speci�ed later in this course.
Proof 3.7.4
To be �lled: To be �lled! p.810 ⇤
Example 3.7.5
To be �lled!
59 S����� 2023
C������ 4
In all this chapter, n is a positive integer and E denotes a vector space on C. Moreover K denotes the
set R of real numbers.
Proposition 4.1.1
In a vector space E which has a spaning family with n elements, every family that has n+1 elements
is not linearly independent.
Proof 4.1.2
To be written [CPY96, Proposition II.37 p.57]. ⇤
Proof 4.1.4
To be written [CPY96, Proposition II.38 p.58]. ⇤
Proposition 4.1.5
Let E be a vector space with �nite dimension n.
60
L����� A������ Lecture Notes C������ 4: Vector Spaces with �nite Dimension
Proof 4.1.6
To be written [CPY96, Proposition II.39 p.58]. ⇤
Remark 4.1.7
According to the previous proposition, if (x1 , x2 , · · · , xp ) is a linearly independent family and if
(y 1 , y 2 , · · · , y q ) is spanning family of E, then p 6 q.
Thus every spanning family has, at least, as many elements as any linearly independent family. We
therefore deduce the two following results.
Proof 4.1.9
to be �lled ⇤
Corollary 4.1.11
Let E be a �nite dimensional vector space of dimension n and let (x1 , x2 , · · · , xp ) a family with p
elements of E.
Example 4.1.12
To be �lled!!!
61 S����� 2023
L����� A������ Lecture Notes C������ 4: Vector Spaces with �nite Dimension
Proof 4.2.2
To be written [CPY96, Theorem II.40 p.59]. ⇤
Proof 4.2.4
To be �lled ⇤
Example 4.2.5
Since the family of vectors (1, X, X 2 , · · · , X n ) is linearly independent, for all n in N, Theorem 4.1.8
allows one to state that the vector space K[X] is in�nite-dimensional.
Proof 4.2.7
To be written [CPY96, Theorem II.42 p.59]. ⇤
Corollary 4.2.8
Let E be a �nite-dimensional vector space with n := dim E.
2. Every linearly independent family of E can be completed so that it becomes a basis of E (in
general there are in�nitely many ways to do so).
Proof 4.2.9
To be �lled! See [CPY96, p.59 Corolaire II.43] ⇤
62 S����� 2023
L����� A������ Lecture Notes C������ 4: Vector Spaces with �nite Dimension
Remark 4.2.10
If E = {0}, the only linearly independent family is the empty family which, in this case, the only
basis of E.
(i) B is a basis of E,
Proof 4.2.12
To be �lled! ⇤
Remark 4.2.13
This fundamental result is of a great use to show that a family B is a basis of a vector space the
dimension of which is known. In most cases we prove that B is a linearly independent family with
n elements.
Example 4.2.14
1. In the R vector space C the family (1, Ê) is linearly independent, i.e. is a basis, if and only if,
Ê is not a real number.
63 S����� 2023
L����� A������ Lecture Notes C������ 4: Vector Spaces with �nite Dimension
Any vector x := (x1 , x2 , x3 , x4 ) of G ful�lls the equality x = (x1 , ≠x1 , ≠x4 , x4 ). This leads
us to think that dim(G) = 2. Let’s prove it by showing that the family of vectors G := (g 1 , g 2 )
de�ned by by setting:
constitutes a basis of G.
3. Since K is a 1-dimensional vector space on K, its vector subspaces can only have a dimension
equal to 0 or 1. Thus K has onmy to vector subspaces that are {0} and K.
Proposition 4.2.17
Any vector subspace F of a �nite dimensional vector space E has, at least one supplementary vector
m
space G. Thus F G = E.
Proof 4.2.18
To be �lled! See [CPY96, p.59 Proposition II.44] ⇤
Remark 4.2.19
Unless F equals {0E } or E, a vector subspace has many supplementary vector subspaces. It is
therefore very important to speak about a supplementary vector subspace.
64 S����� 2023
L����� A������ Lecture Notes C������ 4: Vector Spaces with �nite Dimension
Proposition 4.3.1
Let E be a �nite-dimensional K-vector space of dimension n. A vector space F is isomorphic to E
if and only if F is a �nite-dimensional vector space of dimension n.
Proof 4.3.2
To be �lled! ⇤
Corollary 4.3.3
Any K-vector space of dimension n is isomorphic to Kn .
Proof 4.3.4
To be �lled ⇤
Proof 4.3.6
To be �lled ⇤
Example 4.3.7
Donner l’exemple de Mp,q (R) ƒ Rp ◊ Rq !
Example 4.3.9
Expand [CPY96, p.62 Example II.46]
65 S����� 2023
L����� A������ Lecture Notes C������ 4: Vector Spaces with �nite Dimension
Remark 4.4.2
Let F := (x1 , x2 , · · · , xn ) be a family of vectors of E.
1. Since F is a spanning family of Span F, one can extract a basis of Span F. Thus rk F 6 n
and the equality holds if and only if F is a linearly independent family.
2. If F has r linearly independent vectors, we have rk F 6 r, and the equality case happens
if and only if these r linearly independent vectors span Span F, in other words the equality
happens if and only if any element of F is a linear combination of these r vectors.
f 1 := (1, 0, 0), f 2 := (0, 1, 0), f 3 := (0, 0, 1), f 4 := (1, 1, 1). f 5 := (1, 0, 1).
(i) since (f 1 , f 2 , f 3 ) is a linearly independent family of vectors of R3 , one has the equality
dim(Span(F)) = 3. Thus rk(F) = 3
(ii) since f 1 + f 3 = f 5 the family of vectors G is not linearly independent. Thus, one can write:
dim(Span(G)) < 3. Moreover, since (f 1 , f 3 ) is a linearly independent family of vectors,
one has dim(Span(G)) > 2. One therefore conclude that dim(Span(G)) = 2 and thus, that
rk(G) = 2.
Remark 4.4.5
• If F is a �nite-dimensional space with dimension n, the rank of any linear map u : E æ F is
�nite and smaller or equal to n. moreover, the equality holds if and only if u is surjective.
66 S����� 2023
L����� A������ Lecture Notes C������ 4: Vector Spaces with �nite Dimension
u is injective).
Rank-Nullity Theorem
Remark 4.4.7
In view of De�nition 5.3.12, the previous equality can be rewritten under the form
It is clear that Ker(h) = Span{(1, 1)}. According to the rank nullity theorem we therefore
know that dim(Im(h)) = 2 ≠ 1 = 1. Thus Im(h) = Span{h((1, 0))} = Span{(1, ≠3)}.
(P ) = P (X + 1) ≠ P (X).
From these two points one can deduce that every polynomials with a degree smaller than n is
of the form P (X + 1) ≠ P (X), where deg(P ) 6 n
Corollary 4.4.9
Two �nite-dimensional vector spaces E and F have the same dimension if and only if there exist a
bijective linear map u : E æ F .
Proof 4.4.10
To be �lled! See [CPY96, p.62 Corollaire II.48] ⇤
Remark 4.4.11
Let u : E æ F be a bijective linear map. Its inverse function, denoted u≠1 , is also linear.
67 S����� 2023
L����� A������ Lecture Notes C������ 4: Vector Spaces with �nite Dimension
Proof 4.4.13
To be �lled! See [CPY96, p.62 Corollaire II.50] ⇤
Example 4.4.14
To be written!
Corollary 4.4.15
If u is an endomorphism, of a �nite-dimensional vector space E, it is equivalent to to say that the
map u is injective, or surjective or bijective.
Proof 4.4.16
Obvious in view of the previous Corollary. ⇤
Remark 4.4.17
Be careful if E is not a �nite-dimensional vector space!
• An endomorphism of E may be injective without being surjective, as the example of the fol-
lowing linear map shows:
: R[X] æ R[X]
P ‘æ X · P (X).
Indeed, is clearly linear. Moreover, Ker( ) = {P œ R[X], X · P (X) = 0} = {P :=
qq q+1 s.t. qq
k=0 ak X , q œ N, (a0 , · · · , aq ) œ R
k k+1 = 0}. Thus Ker( ) = {0} and
k=0 ak X
therefore is injective. However P (X) := 1 does not belong to Im( ) which is therefore not
surjective.
• An endomorphism of E may be surjective without being injective, as the example of the fol-
lowing linear map shows:
: R[X] æ R[X]
P ‘æ P Õ (X).
Indeed, any constant polynomial has an image by which equals 0 so is not injective.
q
Besides any polynomial P := qk=0 ak X k of R[X] ful�lls the equality
Ç q å
ÿ ak
X k+1
= P.
k=0
k+1
This makes surjective.
68 S����� 2023
L����� A������ Lecture Notes C������ 4: Vector Spaces with �nite Dimension
69 S����� 2023
C������ 5
Use of Basis
In this chapter, we show how the language of Vector Spaces can be translated into Matrix language
and vice-versa. More precisely, we are considering �nite-dimensional vector spaces and we choose a
basis for each of them. We �rst show that the objects de�ned in Chapters 3 and 4 (vectors spaces and
linear maps) have a matrix representation. Then we study how these representations depend on the
choice of these basis.
The only problem with this kind of notation is that it is leading to a contradiction since the same
quantity, the vector x, has two di�erent values, according to (1.2) and (1.3). However, De�nition 1.1 is
70
L����� A������ Lecture Notes C������ 5: Use of Basis
valid since it does de�ne, in a unique way, vector x. In order to overcome this apparent contradiction,
we must precise, in which basis the coordinates of a given vector are speci�ed.
More generally, de�ne the endomorphism of R3 , denoted u, by setting:
8
< u(e1 ) := 2e1 ≠ 3e2 + 2e3
>
u(e2 ) := ≠3e1 ≠ 6e2 + 6e3
>
:
u(e3 ) := 2e1 ≠ 2e2 + 2e3 .
How can we express simply u(f 1 ), u(f 2 ), u(f 3 ), and, more generally u(y), for any vector y :=
af 1 + bf 2 + cf 3 , where a, b and c are real numbers?
Some other questions arise:
- How the matrix representations of the same endomorphism u, in di�erent basis of E, are linked
to one another?
- Should one use preferably one basis over another, to perform computations?
where the –i are scalar (they are the coordinates or components) of vector x, in the basis e. We write:
â ì
–1
[x]e = .. .
.
–p
In other words, [x]e is the column matrix of the coordinates of vector x, in the basis e. We say that
the column matrix represents the vector x in the basis e.
More generally, if x1 , · · · xq are vectors of E, we denote ([x1 ]e · · · [xq ]e ) or simply [x1 , · · · , xq ]e the
p ◊ q matrix, the j th column of which is [xj ]e .
We easily verify that the map x ‘æ [x]e , from E to Mp,1 (K), is linear, i.e. , for all x and y in E and ⁄
in K, we have [⁄x + y]e = ⁄ [x]e + [y]e . Moreover, since this map transforms the basis e of E into
the standard basis of Mp,1 (K), it is bijective (according to Proposition 3.6.7).
Linear Maps
Let E be a �nite-dimensional vector space with dim(E) = p and let F be a �nite-dimensional vector
space with dim(F ) = q. Let e := (e1 , · · · , ep ) be a basis of E, f := (f 1 , · · · , f q ) be a basis of F and
71 S����� 2023
L����� A������ Lecture Notes C������ 5: Use of Basis
u : E æ F be a linear map. We denote ⁄ij (1 6 i 6 q) the coordinates of the vector u(ej ) in the basis
f , i.e. the scalars which verify:
q
ÿ
u(ej ) = ⁄ij f i . (2.5)
i=1
We denote Mate,f (u) or [u]fe the1 p ◊ q matrix (⁄ij )16i6q,16j6p . It is called the matrix of the linear
map u from basis e to f .
It is clear that the columns of the matrix [u]fe are the column matrices [u(ej )]f . In other words:
The matrix of a linear map u : E æ F , from basis e to f , has, as columns, the column matrices of the
coordinates of the image of the vectors of basis e of E, in the basis f of F :
Ä ä
[u]fe = [u(e1 )]f · · · [u(ep )]f .
It is clear that, if u and v are both linear maps in L(E, F ), we have, for all scalar ⁄,
[⁄u + v]fe = ⁄ [u]fe + [v]fe
Remark 5.2.2
Note that one can also refer to [u]fe as Mate,f (u).
Fondamental Formulae
We will see in this paragraph how the notations we introduced in the previous paragraph are adapted
to the operations already de�ned, from one hand on Matrices, and, from the other hand on linear maps.
Let E be a �nite-dimensional vector space with dim(E) = p, F be a �nite-dimensional vector space
with dim(F ) = q and u : E æ F be a linear map. With the notations introduced at (2.4) and (2.5), we
easily see that, for all vector x in E,
p
! p
ÿ ÿ
u(x) = u –j ej = –j u(ej )
j=1 j=1
p q q p
!
ÿ ÿ ÿ ÿ
= –j ⁄ij f i = ⁄ij –j f i.
j=1 i=1 i=1 j=1
1
The notation [u]fe comes from [CPY96, Chap. III]
72 S����� 2023
L����� A������ Lecture Notes C������ 5: Use of Basis
Conversely, if A is a matrix in Mp,q (K), the map u : X ‘æ AX, from Mp,1 (K) to Mq,1 (K) is a linear
map. Ifone identi�es Mp,1 (K) with Kp and Mq,1 (K) with Kq using standard basis e of Kp and f of
Kq , we see that, for all vector X of Kp , we have the equality:
Proposition 5.2.3
Let E and F be two �nite-dimensional vector spaces with dim(E) = p and dim(F ) = q. Let
e := (e1 , · · · , ep ) be a basis of E, f := (f 1 , · · · , f p ) be a basis of F and u : E æ F be a linear
map. The map u is bijective if, and only if, p = q and if the matrix [u]fe is invertible. In this latter
⇥ ⇤e Ä ä≠1
case, we have the equality: u≠1 f = [u]fe .
Proof 5.2.4
To be �lled! ⇤
73 S����� 2023
L����� A������ Lecture Notes C������ 5: Use of Basis
Theorem 5.2.5
Let E and F be two �nite-dimensional vector spaces, e being a basis of E and f being a basis of F
and let u : E æ F be a linear map. The rank of the matrix [u]fe equals the rank of the linear map u.
In particular it does not depend on the choice of the basis e nor f .
Proof 5.2.6
To be �lled To be �lled! ⇤
Corollary 5.2.7
Let q be a non-zero integer, S := (x1 , · · · , xq ) be a family of vectors of a �nite-dimensional vector
space E and e be a basis of E. The rank of S (which is by de�nition the rank of vector subspace
Span(S) equals the rank of the matrix
([x1 ]e · · · [xq ]e ) .
Proof 5.2.8
To be �lled To be �lled! ⇤
The change of basis matrix, denoted P , from the former basis e to the new one f (or simply the
change of basis matrix P from e to f ) is the matrix the columns of which represent the vectors of
the new basis, given in the coordinate system of the former basis.
⇥ ⇤e
P = [IE ]ef = [f 1 ]e · · · f p . (3.9)
74 S����� 2023
L����� A������ Lecture Notes C������ 5: Use of Basis
Proposition 5.3.2
The change of basis matrix P from the basis e to f . Its inverse P ≠1 is the change of basis matrix
from the basis f to e.
Proof 5.3.3
To be �lled! ⇤
The components of a vector x are modi�ed by a change of basis (from e to f let’s say). Precisely the
relation is:
Let X be the column matrix of the components of a vector in the former basis, Y be the column
matrix of the components of the same vector in the new basis and let P denote the change of basis
matrix. We then have the following equality:
Y = P ≠1 X. (3.10)
75 S����� 2023
L����� A������ Lecture Notes C������ 5: Use of Basis
Example 5.3.6
We keep the data given at Example 5.3.1. Denote x the vector de�ned by x := 2e1 ≠ 3e2 + 4e3 .
Denote also X := [x]e and Y := [x]f . It is clear that
Ö è
2
[x]e = ≠3 .
4
According to (3.10), it is easy to determine the components of the vector x in the basis f . We hence
can write:
Ö è Ö è Ö è
2 ≠4 3 2 28
Y = [x]f = [IE (x)]f = [IE ]fe · [x]e = P ≠1 · X = 0 3 ≠2 · ≠3 = ≠17 .
≠1 6 ≠4 4 ≠36
Every bottom left map cancels with the same map, at the top right. It reminds what happens, in a
di�erent order (i.e. every number at the top right cancels with the bottom left number), with the
Riemann integral of an integrable function over an interval [a, b]. For every c in [a, b]:
⁄ c ⁄ b ⁄ c
⇤
f= f+ f,
a a b⇤
In order to remember Formula (3.11), we introduce below standard conventions (that must be respected
if one does not want to make mistakes). Basis e and f are said to be “former basis” while eÕ and f Õ are
said to be “new basis”.
Let u : E æ F be a linear map. Let e to eÕ be two basis of E and f and f Õ be two basis of F . Let A
be the matrix representation of u from the former basis e and f and let B be the matrix of u from
the new basis eÕ to f Õ .
Denote P the change of basis matrix from e to eÕ and denote Q the change of basis matrix from f
to f Õ . We have the following equality:
76 S����� 2023
L����� A������ Lecture Notes C������ 5: Use of Basis
Example 5.3.8
We keep the linear map de�ned at Example 5.2.1. Hence u : R3 æ R2 , is de�ned, for all (x, y, z) in
R3 , by:
u(x, y, z) := (x + y ≠ 4z, ≠3x + 2y + 5z).
e := (e1 , e2 , e3 ) denotes the standard basis of R3 while f := (f 1 , f 2 ) denotes the standard basis
of R2 . De�ne eÕ := (eÕ1 , eÕ2 , eÕ3 ) and f Õ := (f Õ1 , f Õ2 ) by:
8
< e1 := 2e2 + 3e3 ® Õ
> Õ
f 1 := 2f 1 ≠ 3f 2
e2 := 2e1 ≠ 5e2 ≠ 8e3
Õ &
>
: Õ f Õ2 := 3f 1 ≠ 4f 2
e3 := ≠e1 + 4e2 + 6e3
Let A be the matrix representation of u from the former basis e and f . From Example 5.2.1, we know
that Ç å
1 1 ≠4
A= .
≠3 2 5
We have seen, in Example 5.3.1, that eÕ is a basis of R3 . For the same reason, f Õ is a basis of R2 . If
P denotes the change of basis matrix from e to eÕ and Q the change of basis matrix from eÕ to f Õ ,
then one can write:
Ö è
0 2 ≠1 Ç å
2 3
P = 2 ≠5 4 & Q= .
≠3 ≠4
3 ≠8 6
Let B be the matrix representation of u from the new basis eÕ and f Õ . We have the equality:
Ö è
Ç åÇ å 0 2 ≠1 Ç å
≠4 ≠3 1 1 ≠4 ≠17 52 ≠39
B = Q≠1 AP = 2 ≠5 4 = .
3 2 ≠3 2 5 8 ≠25 19
3 ≠8 6
Proposition 5.3.9
Let A be a square matrix of size p and of rank r. There exist two square invertible matrices, denoted
G and D such that GAD = Jr , where we de�ned:
0 1
1 0 ··· 0
B .. C
B0
B . C
C
B C
B 1 C Ω row r
B. .. C
Jr =B
B ..
C
0 .C
B C
B .. C
@ . A
0 ··· 0
Proof 5.3.10
To be �lled To be �lled! See [CPY96, p.85 Proposition III.5] ⇤
77 S����� 2023
L����� A������ Lecture Notes C������ 5: Use of Basis
Remark 5.3.11
1. Note that Jr can also be denoted by
Ç å
Ir 0p≠r,p≠r
(3.13)
0p≠r,p≠r 0r,r
where 0s,t denotes the null matrix which has s rows and t columns.
2. Multiply a matrix A, from the right hand side (respectively from the left hand side) by a square
and invertible matrix comes to make a change of basis in the departure space (respectively
to make a change of basis in the arrival space) of the map u. It does not modify or a�ect
the rank of the matrix: since A and B both represent the linear map u, we have the equality
rk(A) = rk u = rk(B). Since the rank of Jr is obviously r, we see that the property given in
Proposition 6.3.2 characterizes the matrices with rank r.
3. The way to obtain matrices G and D is to perform Gaussian Elimination (see Chapter 1 for
more details).
Corollary 5.3.13
The rank of a matrix equals the rank of its transpose.
Proof 5.3.14
To be �lled To be �lled! See [CPY96, p.86 Corollaire III.7] ⇤
Corollary 5.3.15
The rank of a p ◊ q matrix equals the rank of the family of its rows vectors in Kq .
Proof 5.3.16
To be �lled! See [CPY96, p.86 Corollaire III.8] ⇤
B = QAP.
Remark 5.3.18
It results from the previous de�nition that all matrices which represent a given linear map u are
equivalent to each other.
78 S����� 2023
L����� A������ Lecture Notes C������ 5: Use of Basis
Example 5.3.19
Matrices A and B, from Example 5.3.8 are equivalent.
Let A be the matrix of an endomorphism u : E æ E in the former basis and let B be the matrix of
the endomorphism u in the new basis. Denote P the change of basis matrix. We have the following
equality:
B = P ≠1 AP. (3.14)
De�nition 5.3.1 (Eigenvalue Space). Let u be an endomorphism of a K vector space E and let ⁄
be an eigenvalue of u. The vector subspace E⁄ := ker(u ≠ ⁄IE ), i.e. the set of all vectors x of E such
that u(x) = ⁄ x, is called the eigenspace of u associated with the eigenvalue ⁄.
Example 5.3.20
Let B := (e1 , e2 , e3 ) be the standard basis of R3 and let u : R3 æ R3 be the linear map de�ned by
setting: 8
< u(e1 ) := ≠e1 + 2e2 + 2e3
>
u(e2 ) := 2e1 ≠ e2 + 2e3
>
:
u(e3 ) := 2e1 + 2e2 ≠ e3 .
De�ne E≠3 := {t œ R3 , f (t) = ≠3t} and E3 := {t œ R3 , f (t) = 3t}. De�ne f 1 := e1 ≠ e3 ,
f 2 := e2 ≠ e3 and f 3 := e1 + e2 + e3 . Denote f := (f 1 , f 2 , f 3 ).
1. Determine the matrix representation of u, from basis e and e, which will be denoted M .
2. Show that f is a basis of R3 .
3. Determine the matrix representation of u, from basis f to f , which will be denoted D.
4. Compute M n , for all n in Nú .
Answers
1. It is clear that the we have the equality:
Ö è
≠1 2 2
M := [u]ee = Mate (u) = 2 ≠1 2 .
2 2 ≠1
2. One just has to solve, in ⁄1 , ⁄2 and ⁄3 , the following linear system:
Ö è Ö è Ö è Ö è
≠1 2 2 0
⁄1 2 + ⁄2 ≠1 + ⁄3 2 = 0 ,
2 2 ≠1 0
79 S����� 2023
L����� A������ Lecture Notes C������ 5: Use of Basis
the solution of which is clearly ⁄1 = ⁄2 = ⁄2 = 0. This proves that the vectors which
constitues the family of vectors f are linearly independent.
4. The equality
D = P ≠1 M P (3.15)
can also be read (by multiplying (3.15) by P from the left-hand side and by P ≠1 from the right-
hand side):
M = P DP ≠1
Therefore, one can write:
M n = P DP ≠1 P DP ≠1 · · · P DP ≠1 = P Dn P ≠1 , (3.16)
and thus
Ö èÖ èÖ 2 è
1 0 1 (≠3)n 0 0 3 ≠ 13 ≠ 13
M =
n
0 1 1 0 (≠3) n
0 ≠ 13 2
3 ≠ 13
1 1 1
≠1 ≠1 1 0 0 3n 3 3 3
Ü ê
2(≠3)n 3n (≠3)n 3n (≠3)n 3n
3 + 3 ≠ 3 + 3 ≠ 3 + 3
(≠3)n 2(≠3)n
= + 33 ≠ (≠3)
n
≠ 3 + 33 + 33
n n n
3 3
.
≠ (≠3) ≠ (≠3) 2(≠3)n
n n
3n 3n 3n
3 + 3 3 + 3 3 + 3
80 S����� 2023
L����� A������ Lecture Notes C������ 5: Use of Basis
Remark 5.3.21
1. In Example 5.3.20, computing An , without using the relation between A and D would be im-
possible.
2. We will see, in Chapter 7, a systematic way to determine the vectors of the family f (of Example
5.3.20), when they exist.
As we will see in Chapter 7, a basis f , as we had in Example 5.3.20, does not necessarily exist. However,
it is still possible to give the matrix representation of a linear map between any basis we choose.
Remark 5.3.23
It results from the previous de�nition that all matrices which represent the same given endomor-
phism u are similar to each other.
Example 5.3.24
Matrices A and D, from Example 5.3.20, are similar.
81 S����� 2023
Part III
82
C������ 6
Ce qu’a fait Gourdon est 100 fois mieux. Regarder comment modi�er ce chapitre de A à Z en prenant
Exemple sur [Gou09, Chap. III.4]
Rajouter des exemples pour chaque section car il n’y en n’a pas
un seul!
In this entire chapter p denotes a positive integer.
6.1 De�nitions
They are many di�erent ways to de�ne the notion of Determinant. Of course they are all equivalent.
In this course we will start from p-linear function to do so.
We identify in the sequel a vector x of Kp with the column matrix [x]e , where e denotes the standard
basis of Kp (in other words one write the components of the vector x in column). In particular it is
the same thing to consider a family (x1 , x2 , · · · , xp ) of p vectors of Kp or to consider the p ◊ p matrix
A := ([x1 ]e · · · [xp ]e ), the columns of which are the column matrices [xj ]e .
1. Whatever the index i, when one freezes all the columns of a matrix
A := ([x1 ]e · · · [xi ]e · · · [xp ]e ), except for the ith , one gets a map Xi ‘æ
b ([x1 ] · · · [xi≠1 ] Xi [xi+1 ] · · · [xp ] ) from K to K which is linear.
e e e e p
2. If the matrix A has two equal columns (Xi = Xj for two di�erent indices i and j) then b(A) =
0.
Proposition 6.1.2
Let b be an alternating p-linear form on Kp and B be the matrix obtained by exchanging two columns
of Matrix A. One has the equality b(B) = ≠b(A).
Proof 6.1.3
To be �lled To be �lled! See [CPY96, p.106 Proposition IV.13] ⇤
83
L����� A������ Lecture Notes C������ 6: A First Look at Determinants
Proposition 6.1.4
An alternating p-linear form on Kp is completely determined by the value of b(Ip ), where Ip denotes
the identity matrix of size p.
Proof 6.1.5
To be �lled To be �lled! See [CPY96, p.106 Proposition IV.14] ⇤
Remark 6.1.6
It is easy to verify that the set B of the alternating p-linear forms on Kp is a vector subspace of the
vector space KMp (K) . The previous proposition shows that two alternating p-linear forms on Kp are
collinear. In other words, the vector space B has a dimension not greater than 1. In order to show
that dim(B) is 1 exactly, one just needs to �nd an alternating p-linear forms on Kp di�erent from 0.
where A := (ai,j )(i,j)œJ1,pK belongs to Mp (K), ‡p denotes the set of all permutations 1 of {1, · · · , p}
and sgn(‡) denotes the signature of the permutation ‡. In (1.1), the sum is computed over all permu-
tations ‡ of the set {1, 2, ..., p}.
Remark 6.1.7
Note that (1.1) also reads
ÿ p
Ÿ
b(A) := sgn(‡) a‡(i),i · b(Ip ). (1.2)
‡œ ‡p i=1
One can check that the form b, de�ned in (1.1), for which we impose that b(Ip ) := 1, is a non zero
alternating linear form. This is this particular linear form we will be interested in in the sequel.
Remark 6.1.9
1. Gathering both Equality (1.2) and De�nition 6.1.8 we obtain the Leibniz formula, which reads
for an p ◊ p matrix A:
ÿ p
Ÿ
det(A) := sgn(‡) a‡(i),i . (1.3)
‡œ ‡p i=1
1
A permutation is a function that reorders this set of integers (here the set is 1, 2, ..., p). The value in the ith position after
the reordering ‡ is denoted by ‡(i). For example, for p = 3, the original sequence 1, 2, 3 might be reordered to ‡ := (2, 3, 1)
(i.e. ‡(1) = 2, ‡(2) = 3 and ‡(3) = 1). The set of all such permutations (also known as the symmetric group on p elements)
is denoted by ‡p . For each permutation ‡, sgn(‡) denotes the signature of ‡, a value that is +1 whenever the reordering
given by ‡ can be achieved by successively interchanging two entries an even number of times, and ≠1 whenever it can be
achieved by an odd number of such interchanges.
84 S����� 2023
L����� A������ Lecture Notes C������ 6: A First Look at Determinants
to denote det(A).
Theorem 6.2.1
Determinant have the following properties.
4. The value of a determinant will not change if one adds, to one of its columns a linear combi-
nation of its other columns.
5. If the family of column matrices A1 , · · · , Ap is not linearly independent (i.e. the corresponding
family of vectors of Kp is not linearly independent) then det(A1 , · · · Ap ) = 0.
Proof 6.2.2
To be �lled! See [CPY96, p.108 Proposition IV.17] ⇤
A direct consequence of the p-linearity of the determinant is that, for all ⁄ in K and A in Mp (K), we
have:
det(⁄A) = ⁄p det(A). (2.4)
Theorem 6.2.3
Let A and B be two matrices in Mp (K). We have the equality:
Proof 6.2.4
To be �lled! See [CPY96, p.109 Theorème IV.18] ⇤
Theorem 6.2.5
In order the p ◊ p matrix A is invertible it is necessary and su�cient that its determinant is non zero.
In this latter case, one has the equality:
1
det(A≠1 ) = . (2.6)
det(A)
85 S����� 2023
L����� A������ Lecture Notes C������ 6: A First Look at Determinants
Proof 6.2.6
To be �lled! See [CPY96, p.109 Theorème IV.19] ⇤
Theorem 6.2.7
For every matrix A of Mp (K), the following equality holds:
det(tA) = det(A).
Proof 6.2.8
To be �lled To be �lled! See [CPY96, p.109 Theorème IV.19] ⇤
Theorem 6.2.9
Determinant have the following properties.
4. The value of a determinant will not change if one adds, to one of its rows a linear combination
of its other rows.
5. If the rows of matrices of a square matrix A are not linearly independent (i.e. the corresponding
family of vectors of Kp is not linearly independent) then det(A) = 0.
Proof 6.2.10
To be �lled One applies Theorem to the matrix tA since the columns of this latter are nothing but the
rows of A. ⇤
De�nition 6.3.1
Let A := (aij ) be a matrix of Mp (K). We call minor of the coe�cient aij in A, and we denote it
Mij (A), the determinant of the matrix of Mp≠1 (K) obtained by removing the ith row and the j th
column of A.
Proposition 6.3.2
We have the equality:
p
ÿ
det(A) = aij (≠1)i+j Mij (A).
j=1
Proof 6.3.3
To be �lled! See [CPY96, p.110 Proposition IV.23]
86 S����� 2023
L����� A������ Lecture Notes C������ 6: A First Look at Determinants
Theorem 6.3.6
Let A := (aij )16i,j6p be a matrix of Mp (K). We have the following equalities:
p
ÿ
det(A) = aij Cij (A) (expanding along the ith row),
j=1
ÿp
det(A) = aij Cij (A) (expanding along the j th column).
i=1
Proof 6.3.7
To be �lled To be �lled! See [CPY96, p.112 Theoreme IV.25] ⇤
Corollary 6.3.8
If det(A) ”= 0, then
1
A≠1 = t
(Ã)
det(A)
where à is the matrix of cofactors i.e. Ãij = Cij (A), for every (i, j) in J1, pK.
Proof 6.3.9
To be �lled To be �lled! See [CPY96, p.112 Corollaire IV.26] ⇤
In view of Theorem ??, it is clear that the determinant of an upper triangular matrix equals the
product of its diagonal terms.
87 S����� 2023
L����� A������ Lecture Notes C������ 6: A First Look at Determinants
• Expansion
The general method (i.e. the one described above) gives good results for numerical matrices (i.e.
matrices with numbers instead of letters). However one can often improve it using the expansion
along a row or a column when not many non zero coe�cients left in this very row or column.
When one wants to compute a determinant, denoted let’s say p which has a particular form
(with p rows and p columns and p being unspeci�ed), one can use this way of computing deter-
minant in order to get a recursive formula which links p and p≠1 .
• Numerical Techniques
In order to compute a determinant using a computer, we use methods adapted to each speci�c
case. The LU factorization is a good example of such a method.
Corollary 6.4.2
The rank of a matrix is the greatest order of an extracted determinant di�erent from 0.
Proof 6.4.3
To be �lled To be �lled! See [CPY96, p.114 Corollaire IV.30] ⇤
Theorem 6.4.4
The rank of a matrix equals the rank of its transpose.
Proof 6.4.5
To be �lled! See [CPY96, p.115 Corollary IV.31] ⇤
88 S����� 2023
Part IV
89
C������ 7
Diagonalization
u(x) = ⁄ x.
Remark 7.1.2
Since, whatever the scalar ⁄, we have u(0) = 0 = ⁄ 0, the requirement x ”= 0 is essential in the
previous de�nition.
90
L����� A������ Lecture Notes C������ 7: Diagonalization
Thus w belongs to E1 .
2. From the previous question we deduce that both u and v are eigenvectors associated to the
eigenvalue ≠1. Moreover E≠1 is the eigenspace associated to the eigenvalue ≠1.
Besides, and that w is an eigenvector associated to the eigenvalue 1. and E1 is the eigenspace
associated to the eigenvalue 1.
Proposition 7.1.5
Let E be a �nite-dimensional K-vector space with dim(E) = p and let u be an endomorphism of
E. The scalar det([u]ee ) does not depend on the basis e of E. We denote it det(u) and is called the
determinant of the endomorphism u.
Proof 7.1.6
To be �lled To be �lled! See [CPY96, p.136 Proposition V.3] ⇤
Proposition 7.1.7
Let E be a K-�nite-dimensional vector space with dim(E) = p, u an endomorphism of E and ⁄ be
an element of K. The following conditions are equivalent.
1. ⁄ is an eigenvalue of u.
Proof 7.1.8
To be �lled To be �lled! See [CPY96, p.136 Proposition V.4] ⇤
91 S����� 2023
L����� A������ Lecture Notes C������ 7: Diagonalization
is a polynomial in X. Its degree equals p = dim(E). This polynomial is called the Characteristic
polynomial.
≠1 ≠ 3X 2 2
1
‰u (X) = det( · (3A ≠ 3XI3 )) = 3≠3 det(·(3A ≠ 3XI3 )) = 2 ≠1 ≠ 3X 2
3
2 2 ≠1 ≠ 3X
Å ã3 3 ≠ 3X 2 2 1 2 2
1 (3 ≠ 3X)
= 3 ≠ 3X ≠1 ≠ 3X 2 = 1 ≠1 ≠ 3X 2
3 33
3 ≠ 3X 2 ≠1 ≠ 3X 1 2 ≠1 ≠ 3X
1 2 2
(3 ≠ 3X) (3 ≠ 3X)
= 0 ≠3 ≠ 3X 0 = · (3 + 3X)2 = (1 ≠ X) · (1 + X)2 .
33 33
0 0 ≠3 ≠ 3X
92 S����� 2023
L����� A������ Lecture Notes C������ 7: Diagonalization
Proposition 7.2.2
Let E be a K-�nite-dimensional vector space with dim(E) = p, u be an endomorphism of E and
let f := (f 1 , · · · , f p ) be a basis of E. The matrix [u]ff is diagonal if, and only if, the vectors f i
are eigenvectors of u. The ith coe�cient of its diagonal is then the eigenvalue of u associated to the
corresponding eigenvector f i .
Proof 7.2.3
To be �lled To be �lled! See [CPY96, p.138 Proposition V.8] ⇤
Proposition 7.2.4
If an endomorphism u of a K-�nite-dimensional vector space E, with dim(E) = p is diagonalizable,
its characteristic polynomial ‰u splits over K (there are p roots of ‰u in K. Each of these roots is
counted which its multiplicity).
Proof 7.2.5
To be �lled To be �lled! See [CPY96, p.139 Proposition V.9] ⇤
Lemma 7.2.6
Let u be an endomorphism of a K-�nite-dimensional vector space E, with dim(E) = p and
⁄1 , · · · , ⁄m some distinct eigenvalues of u and x1 , x2 , · · · , xm eigenvectors of u associated to
⁄1 , · · · , ⁄m such that u(xi ) = ⁄i xi (i.e. xi œ ker(u ≠ ⁄i IE ).
If x1 + · · · + xm = 0 then x1 = · · · = xm = 0.
Proof 7.2.7
To be �lled To be �lled! See [CPY96, p.139 Proposition V.10] ⇤
Remark 7.2.8
The previous result means that, for distinct eigenvalues ⁄1 , · · · , ⁄m , the corresponding eigenspaces
E⁄i := ker(u ≠ ⁄i IE ) all are in direct sum.
Theorem 7.2.9
Let u be an endomorphism of a K-�nite-dimensional vector space E, with dim(E) = p. If the char-
acteristic polynomial of u has p distinct roots in K, i.e. if u has p eigenvalues, then u is diagonalizable.
Proof 7.2.10
To be �lled! See [CPY96, p.140 Theorème V.12]
⇤
Proposition 7.2.11
Let u be an endomorphism of a K-�nite-dimensional vector space E and let ⁄ be an eigenvalue of
u. Denote r the multiplicity order of ⁄ as a root of ‰u . We have the following inequality:
dim(ker(u ≠ ⁄IE )) 6 r.
93 S����� 2023
L����� A������ Lecture Notes C������ 7: Diagonalization
Proof 7.2.12
To be �lled! See [CPY96, p.140 Proposition V.13] ⇤
1. The characteristic polynomial ‰u splits in K (i.e. it factors completely in K) and for all i in
J1, mK, the equality dim(ker(u ≠ ⁄i IdE )) = ri holds.
m
ÿ
2. dim(ker(u ≠ ⁄i IdE )) = p. In other words, one can write:
i=1
m
ÿ
dim(E⁄i ) = dim(E) = p.
i=1
In this case, one can get a basis of E by gathering all the basis of every eigenspace ker(u ≠ ⁄i IdE ).
Proof 7.2.14
To be �lled! See [CPY96, p.140 Proposition V.14]
⇤
We will keep, for matrices, the terminology we used for endomorphisms. We therefore call eigenvalue
(resp. eigenvector) of A every eigenvalue (resp. every eigenvector) of the associated endomorphism u.
The kernel of u ≠ ⁄IE , denoted ker(A ≠ ⁄Ip ), is called eigenspace of A, associated to the eigenvalue
⁄, and ‰A de�ned by setting ‰A (X) := det(A ≠ XIp ), is called the characteristic polynomial of A.
Theorem 7.3.2
A matrix A of Mp (K) is diagonalizable if, and only if there exists in Mp (K), an invertible matrix,
denoted P , and a diagonal matrix, denoted D, such that:
D = P ≠1 AP. (3.1)
94 S����� 2023
L����� A������ Lecture Notes C������ 7: Diagonalization
Proof 7.3.3
To be �lled To be �lled! See [CPY96, p.141 Theorème V.16] ⇤
Remark 7.3.4
Matrices A and D are similar. Thus a matrix is diagonalizable if, and only if, it is similar to a diagonal
matrix.
1. The characteristic polynomial ‰A splits in K (i.e. it factors completely in K) and for all i in
J1, mK, the equality dim(ker(A ≠ ⁄i Ip )) = ri holds.
m
ÿ
2. dim(ker(A ≠ ⁄i Ip )) = p. In other words, one can write:
i=1
m
ÿ
dim(E⁄i ) = dim(Kp ) = p.
i=1
In this case, one can get a basis of Kp by gathering all the basis of every eigenspace ker(A ≠ ⁄i Ip ).
Proof 7.3.6
One just has to apply Proposition 7.2.13 with the endomorphism
u : Kp æ Kp
X ‘æ AX
Diagonalization in practice
Whenever A in Mp (K), is diagonalizable, diagonalize it means �nding an invertible matrix, denoted
P , and a diagonal matrix, denoted D such that the following equality holds:
D = P ≠1 AP
The matrix P is the change of basis matrix from the standard basis e (also called the former basis) to
a new basis, denoted f , which is constituted of the eigenvectors of A. Matrix D is the repsentation
matrix of the endomorphism u, from basis f to itself. In other words, D = [u]ff .
Starting from matrix A, we get matrices D and P by the following manner.
95 S����� 2023
L����� A������ Lecture Notes C������ 7: Diagonalization
1. The eigenvalues of the matrix A are the roots of the characteristic polynomial ‰A (where ‰A (X) =
det(A ≠ XIp )).
2. We determine the eigenspaces of A and we choose a basis f := (f 1 , · · · , f p ) of Kp constituted
of eigenvectors of A (we obtain it by taking the union of the basis of the eigenspaces). The ith
column of the matrix P is then constituted of the components of f i in the standard basis of Kp .
3. We have D = diag(⁄1 , ·, ⁄p ) where the ⁄i are the eigenvectors of the matrix A. Each and every
one of them appear on the diagonal of D, a number of times which equals its multiplicity, as a
root of ‰A . They are ordered such that f i is associated to ⁄i .
3≠X 4
‰u (X) = = (3 ≠ X)(2 ≠ X) ≠ 20 = X 2 ≠ 5X ≠ 14.
5 2≠X
It is easy to determine that the roots of ‰u are ≠2 and 7. Thus one can write:
2. We are now looking for the eigenvectors of A. By de�nition, we know that the two eigenvectors
are: E≠2 := Ker(A + 2I2 ) and E7 := Ker(A ≠ 7I2 ).
One then concludes that E≠2 = {(x, ≠ 54 x), x œ R} = SpanR {(≠4, 5)}. We can choose
x1 := (≠4, 5) as an eigenvector of A, associated to the eigenvalue ≠2. Note moreover
that (≠4, 5) is a basis of E≠2 .
• Make E7 explicit
Let (x, y) be in R2 .
! ! !
≠4 4 x 0
(x, y) œ E7 ≈∆
t
· = ≈∆ ≠4x + 4y = 0 ≈∆ y = x
5 ≠5 y 0
One then concludes that E7 = {(x, x), x œ R} = SpanR {(1, 1)}. We can choose
x2 := t(1, 1) as an eigenvector of A, associated to the eigenvalue 7. Note moreover that
(1, 1) is a basis of E7 .
96 S����� 2023
L����� A������ Lecture Notes C������ 7: Diagonalization
Finally gathering the two basis of the eigenspaces, we know that B := (x1 , x2 ) is a basis of
R2 , constituted only with eigenvectors of A.
Besides, denote e := (e1 , e2 ) the standard basis R2 and let u : R2 æ R2 be the linear map,
the matrix representation (from e to e) of which is A. In other words, one has the equality:
A = [u]ee .
De�ne D := [u]B B . By de�nition of the xi , we know that u(x1 ) = ≠2x1 and that u(x2 ) = 7x2 .
It is therefore obvious that
u(x1 ) u(x2 )
Ç å
D = [u]B = ≠2 0 x1 .
B
0 7 x2
Moreover, denote P the change of basis matrix from e and B. By de�nition of P , one has the
equality:
x x2
Ç 1 å
P = [u]B =
e ≠4 1 e1 .
5 1 e2
D = [u]B
B = [u]e [u]e [u]B = P
B e e ≠1
AP. (3.2)
One can here verify that Equality holds (3.2) by �rst computing P ≠1 and then by computing
P ≠1 AP .
We easily get !
≠1/9 1/9
P ≠1 =
5/9 4/9
and then
! Ç å ! !
≠1/9 1/9 3 4 ≠4 1 ≠2 0
P ≠1 AP = · · =
5/9 4/9 5 2 5 1 0 7
which is nothing but D. We therefore recover the result which states that Equality (3.2) holds.
7.4 Applications
Computation of the nth power of a matrix
A = P DP ≠1 . (4.3)
It is therefore easy to deduce, that, for every n in N,
An = P DP ≠1 P DP ≠1 · · · P DP ≠1 = P Dn P ≠1 . (4.4)
97 S����� 2023
L����� A������ Lecture Notes C������ 7: Diagonalization
Matrix D (resp. A) represents the endomorphism u in the basis f , which is constituted with eigenvec-
tors of u (resp. in the standard basis e). Equality (4.4) expresses simply the fact that both matrices An
and Dn represent the same endomorphism (namely un ), respectively in the basis e and in the basis f .
Note moreover that the product of two diagonal matrices is very easy to compute. indeed, the product
matrix is nothing but a diagonal matrix, the ith term of which is the product of the ith terms of both
matrices. We therefore deduce that the diagonal matrix Dn is obtained by taking the nth power of each
and every of the factors on the diagonal of D.
Many other applications of diagonalization will be seen in Exercises. Note however that Diagonaliza-
tion can also be successfully used to solve di�erential linear system with constant coe�cients, to study
system of Constant-recursive sequences.
J’en suis à [CPY96, p.143 Applications]
98 S����� 2023
Part V
99
C������ 8
Isometries
100
C������ 9
Euclidean Spaces
101
Part VI
102
L����� A������ Lecture Notes C������ 8: Euclidean Spaces
1. PCA (ACP)
3. Exponentielles de matrices
104
Appendix A
Example A.1.2
Let E := {a, b, c, d}, F := {1, 2, 3, 4} and let F Õ be the set de�ned by setting F Õ := {1, 2, 3, 4, 5}.
De�ne the maps f, g and h by setting:
h : E æ F
f : E æ F g : E æ F
a ‘ æ 4
a ‘ æ 4 a ‘ æ 2
a ‘ æ 1
b ‘ æ 2 , b ‘ æ 2 & .
b ‘ æ 2
c ‘ æ 1 c ‘ æ 1
c ‘ æ 1
d ‘ æ 3 d ‘ æ 3
d ‘ æ 3
105
L����� A������ Lecture Notes C���.A M��� & A�����������
E F E F E F
f g h
a 1 a 1 a 1
b 2 b 2 b 2
c 3 c 3 c 3
d 4 d 4 d 4
Both maps f and g are correctly de�ned. However, h is not correctly de�ned since h(a) = 1 and
h(a) = 4.
Let us also consider the maps s, i and k, from E to F Õ , de�ned by their graph, given below:
E FÕ E FÕ E FÕ
s i k
a 1 a 1 a 1
b 2 b 2 b 2
5 5 5
c 3 c 3 c 3
d 4 d 4 d 4
Both maps s and i are correctly de�ned. However, k is not correctly de�ned since k(a) = 1 and
k(a) = 4.
Example A.1.3
Let g, h, Ï and  be the maps de�ned by setting:
g : R æ R h : R æ R≠ ú Ï : R+ æ R+ Â : R æ R+
x ‘æ x2 x ‘æ x2 x ‘æ x2 x ‘æ x2
It is clear that the maps g, Ï and  are correctly de�ned while h is not a map since it is impossible
for the square of a real number to be negative.
Example A.1.4
Let us also consider the map · : E æ F , the graphic representation of which is given below:
E F
· B
a 1
b 2
c 3
d 4
Remark A.1.5
From these previous de�nitions, one easily sees that the range of function is noting but the image of
its domain. In other words, for any map f : E æ F , one has: range(f ) = f (E).
Remark A.2.2
An alternative characterization of injectivity, which is an alternative to (2.4) is:
Example A.2.3
In Example A.1.2 above, only f and i are injective functions.
Example A.2.4
If one considers the maps g, h, Ï and  de�ned in Example A.1.3, it is easy to see that:
Since both a and b are positive, the case a = ≠b is impossible and thus we conclude that
Example A.2.6
In Example A.1.2 above, only f is a surjective functions.
Example A.2.7
If one still considers the maps g, h, Ï and  de�ned in Example A.1.3, it is easy to see that:
• g is not surjective since, e.g. ≠4, does not have a preimage. Indeed, ≠4 = x2 has no real
solution. For the exact same reason h is not surjective either.
Ô
• Both
Ô Ï and  are surjective since all b in R+ has a preimage in R+ . Indeed, Â( b) = b =
Ï( b), for all b in R+ .
Theorem A.2.8
Let f : E æ F be a map. The two following statements are equivalent:
1. f is surjective
2. f (E) = F
Proof A.2.9
To be �lled! ⇤
’y œ F, f (f ≠1 (y)) = y. (2.8)
We usually call f ≠1 the inverse function of f .
Example A.2.11
In Example A.1.2 above, only f is a bijective function.
Example A.2.12
If one still considers the maps g, h, Ï and  de�ned in Example A.1.3, it is clear, in view of Examples
A.2.4 and A.2.7, that only Ï is bijective. Its inverse is the map Ï≠1 : R+ æ R+ de�ned by: Ï≠1 (u) :=
Ô
u, for every u in R+ .
Remark A.2.13
We have to be extremely cautious with the notation f ≠1 . Indeed, this notation can be used in two
very di�erent situations. Either one wants to speak about:
• the subset f ≠1 (B) of E, where B is a subset of F (as de�ned in (1.3)). Such a set always exists
(i.e. we do not need f to be bijective).
Hence, in example A.1.4, we made explicit the set · ≠1 (B) while · is not bijective and therefore · ≠1
(i.e. the inverse of · ) does not exist.
Example A.2.14
Let f : E æ F be an application. Show that:
(i) f is injective if and only if ’A µ E, A = f ≠1 (f (A)).
(ii) f is surjective if and only if ’C µ F, f (f ≠1 ((C)) = C.
(iii) f is bijective if and only if:
a) ’A µ E, A = f ≠1 (f (A)),
b) ’C µ F, f (f ≠1 ((C)) = C.
Let A be a subset of E and C be a subset of F . By de�nition we have:
[CPY96] Gilles Christol, Philippe Pilibossian, and Sleiman Yammine. Algèbre 2 - Espaces vectoriels,
matrices, systèmes linéaires. Ellipses, 1996.
[Gou09] Xavier Gourdon. Les maths en tête. Algèbre - 2e édition. Les maths en tête. Ellipses, 2009.
[HJ13] Roger A. Horn and Charles R. Johnson. Matrix Analysis. Cambridge University Press, Cam-
bridge, second edition, 2013.
[Str16] Gilbert Strang. Introduction to Linear Algebra. Wellesley-Cambridge Press; 5th edition, 2016.
111
Index
Matrices
Similar matrices, 81
Matrix
Change of basis -, 75
Equivalent matrices, 79
nilpotent, 7
Rank
-Nullity Theorem, 68
of a family of vectors, 67
of a linear map, 67
of a matrix, 17
Theorem
Rank-Nullity, 68
112
List of Symbols
K: denotes either the set of real numbers R or the set of complex numbers C, 2
J1, pK: intervall of all integers from 1 to p, 1 and p being both included, 2
113