Linear Algebra
Linear Algebra
LINEAR ALGEBRA
and its
APPLICATIONS
M.Thamban Nair
Department of Mathematics
Indian Institute of Technology Madras
Contents
Preface iv
1 Vector Spaces 1
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Definition and Some Basic Properties . . . . . . . . . . 2
1.3 Examples of Vector Spaces . . . . . . . . . . . . . . . . 4
1.4 Subspace and Span . . . . . . . . . . . . . . . . . . . . 8
1.4.1 Subspace . . . . . . . . . . . . . . . . . . . . . 8
1.4.2 Linear Combination and Span . . . . . . . . . . 11
1.5 Basis and Dimension . . . . . . . . . . . . . . . . . . . 14
1.5.1 Dimension of a Vector Space . . . . . . . . . . 18
1.5.2 Dimension of Sum of Subspaces . . . . . . . . . 22
2 Linear Transformations 25
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . 25
2.2 What is a Linear Transformation? . . . . . . . . . . . 26
2.3 Space of Linear Transformations . . . . . . . . . . . . 31
2.4 Matrix Representations . . . . . . . . . . . . . . . . . 32
2.5 Rank and Nullity . . . . . . . . . . . . . . . . . . . . . 35
2.6 Composition of Linear of Transformations . . . . . . . 38
2.7 Eigenvalues and Eigenvectors . . . . . . . . . . . . . . 40
2.7.1 Definition and examples . . . . . . . . . . . . . 40
2.7.2 Existence of an eigenvalue . . . . . . . . . . . . 43
2.7.3 Diagonalizability . . . . . . . . . . . . . . . . . 45
ii
Contents iii
iv
1
Vector Spaces
1.1 Introduction
The notion of a vector space is an abstraction of the familiar set of
vectors in two or three dimensional Euclidian space. For example, let
~x = (x1 , x2 ) and ~y = (y1 , y2 ) be two vectors in the plane R2 . Then
we have the notion of addition of these vectors so as to get a new
vector denoted by ~x + ~y , and it is defined by
~x + ~y = (x1 + y1 , x2 + y2 ).
This addition has an obvious geometric meaning: If O is the coordi-
nate origin, and if P and Q are points in R2 representing the vectors
~x and ~y respectively, then the vector ~x + ~y is represented by a point
R in such way that OR is the diagonal of the parallelogram for which
OP and OQ are adjacent sides.
Also, if α is a positive real number, then the multiplication of ~x
by α is defined by
α~x = (αx1 , αx2 ).
Geometrically, the vector αx~ is an elongated or contracted form of
~x in the direction of ~x. Similarly, we can define α~x with a nega-
tive real number α, so that α~x represents in the negative direction.
Representing the coordinate-origin by ~0, and −~x := (−1)~x, we see
that
~x + ~0 = ~x, ~x + (−~x) = ~0.
We may denote the sum ~x + (−~y ) by ~x − ~y .
Now, abstracting the above properties of vectors in the plane, we
define the notion of a vector space.
We shall denote by F the field of real numbers or the field of
complex numbers. If special emphasis is required, then the fields
of real numbers and complex numbers will be denoted by R and C,
respectively.
1
2 Vector Spaces
x + θ1 = x and x + θ2 = x ∀x ∈ X.
θ2 = θ2 + θ1 = θ1 + θ 2 = θ 1 .
x + x0 = θ and x + x00 = θ.
−x := x̃.
0x = (0 + 0)x = 0x + 0x,
we have 0x = θ. Now,
x + (−1)x = [1 + (−1)]x = 0x = θ
x − y := x + (−y).
4 Vector Spaces
Fn := {(α1 , . . . , αn ) : αi ∈ F, i = 1, . . . , n}.
x = a0 + a1 t + . . . + an tn
For x = a0 + a1 t + . . . an tn , y = b0 + b1 t + . . . + bn tn in Pn and
α ∈ F,
−x = −a0 − a1 t − . . . − an tn .
♦
EXAMPLE 1.3 (Space P) Let P be the set of all polynomials
with coefficients in F, i.e., x ∈ P if and only if x ∈ Pn for some
n ∈ {0, 1, 2, . . .}. For x, y ∈ P, let n, m be such that x ∈ Pn and
y ∈ Pm . Then we have x, y ∈ Pk , where k = max {n, m}. Hence we
can define x + y and αx for α ∈ F as in Pk . With this addition and
scalar multiplication, it follows that P is a vector space. ♦
EXAMPLE 1.4 (Space Fm×n ) Let V = Fm×n be the set of all
m × n matrices with entries in F. If A is a matrix with its ij-th
entry aij , then we shall write A = [aij ]. It is seen that V is a vector
space with respect to the addition and scalar multiplication defined
as follows: For A = [aij ], B = [bij ] in V , and α ∈ F,
In this space, −A = [−aij ], and the matrix with all its entries are
zeroes is the zero element. ♦
EXAMPLE 1.5 (Space Fk ) This example is a special case of the
last one, namely,
Fk := Fk×1 .
This vector space is in one-one correspondence with Fk . One such
correspondence is given by F : Fk → Fk defined by
x1
x2
F ((x1 , . . . , xk )) =
. . . ,
(x1 , . . . , xk ) ∈ Fk .
xk
♦
6 Vector Spaces
EXAMPLE 1.8 (Space R[a, b]) Let R[a, b] be the set of all real
valued Riemann integrable functions on [a, b]. From the theory of
Riemann integration, it follows that if x, y ∈ R[a, b] and α ∈ F, then
x + y and αx defined pointwise belongs to R[a, b]. It is seen that
(Verify) R[a, b] is a vector space over R. ♦
V = V1 × · · · × Vn ,
♦
Exercise 1.5 Let Ω be a nonempty set and W be a vector space.
Let F(Ω, W ) be the set of all functions from Ω into W . For f, g ∈
F(Ω, W ) and α ∈ F, let F + G and αF be defined point-wise, i.e.,
Show that F(Ω, W ) is a vector space over F with the above opera-
tions. ♦
Exercise 1.6 In Exercise 1.5, let Ω = {1, . . . , n} and W = F. Show
that the map T : F(S, F) → Fn defined by
is bijective.
Also show that for every f, g in F(Ω, F) and α ∈ F,
T (f + g) = T (f ) + T (g), T (αf ) = αT (f ).
θ = 0 x ∈ V0 and − x = (−1)x ∈ V0 .
Subspace and Span 9
Thus, conditions (c) and (d) in the definition of a vector space are
satisfied for V0 . All the remaining conditions can be easily verified
as elements of V0 are elements of V as well.
V0 = {(x1 , . . . , xn ) ∈ Fn : a1 x1 + . . . + an xn = 0}.
Then V0 is a subspace of Fn .
Recall from school geometry that, for F = R and n = 3, the
subspace
V0 = {(x1 , x2 , x3 ) ∈ R3 : a1 x1 + +a2 x2 + a3 x3 = 0}
Exercise 1.9 Let P[a, b] be the vector space P over R taking its
elements as continuous real valued functions defined on [a, b]. Then
the space P[a, b] is a subspace of C k [a, b] for every k ≥ 1. ♦
Exercise 1.14 Prove that the only proper subspaces of R2 are the
straight lines passing through the origin. ♦
span {u1 , . . . , un }.
Let
a11 a12 a1n
a21 a22 a2n
. . . , u2 := . . . , . . . , un := . . . .
u1 :=
Exercise 1.20 Let uj (t) = tj−1 , j ∈ N. Show that span of {u1 , . . . , un+1 }
is Pn , and span of {u1 , u2 , . . .} is P. ♦
x + E := {x + u : u ∈ E},
E1 + E2 := {x1 + x2 : x1 ∈ E1 , x2 ∈ E2 }.
The set E1 + E2 is called the sum of the subsets E1 and E2 .
V1 + V2 = span (V1 ∪ V2 ).
V1 + V2 ⊆ span (V1 ∪ V2 ) ⊆ V1 + V2 ,
α1 u1 + · · · + αn xn = 0 =⇒ αi = 0 ∀ i = 1, . . . , n.
♦
If {u1 , . . . , un } is a linearly independent (respectively, dependent)
subset of a vector space V , then we may also say that u1 , . . . , un are
linearly independent (respectively, dependent) in V .
Note that a linearly dependent set cannot be empty. In other
words, the empty set is linearly independent!
Basis and Dimension 15
Exercise 1.28 For λ ∈ [a, b], let uλ (t) = exp (λt), t ∈ [a, b]. Show
that {uλ : λ ∈ [a, b]} is an uncountable linearly independent subset
of C[a, b]. ♦
Exercise 1.29 If {u1 , . . . , un } is a basis of a vector space V , then
show that every x ∈ V , can be expressed uniquely as
x = α1 u1 + · · · + αn un ,
Theorem 1.8 If a vector space has a finite spanning set, then it has
a finite basis. In fact, if S is a finite spanning set of V , then there
exists a basis E ⊆ S.
span S1 = span S = V.
Basis and Dimension 19
x1 = α1 u1 + · · · + αn un .
(n) (n)
xn = α1 x1 + · · · + αn−1 xn−1 + αn(n) un .
(n)
Since {x1 , . . . , xn } is linearly independent, it follows that αn 6= 0.
Hence,
un ∈ span {x1 , . . . , xn }.
Consequently,
x := α1 u1 + . . . + αk uk = −(β1 v1 + . . . + β` v` ) ∈ V1 ∩ V2 = {0}
For the proof of the above theorem we shall make use of the
following result.
Then define
(
E1 if v2 ∈ span (E1 ),
E2 =
E1 ∪ {v2 } if v2 ∈
6 span (E1 ).
Then
k
X `
X m
X
x := αi ui + β i vi = − γ i wi ∈ V 1 ∩ V 2 .
i=1 i=1 i=1
2.1 Introduction
We may recall from the theory of matrices that if A is an m × n
matrix, and if x is an n-vector, then Ax is an m-vector. Moreover,
for any two n-vectors x and y, and for every scalar α,
d d d d d
(f + g) = f + g, (αf ) = α f.
dt dt dt dt dt
Z b Z b Z b Z b Z b
(f +g)(t)dt = f (t) dt+ g(t) dt, (αf )(t) = α f (t) dt,
a a a a a
Z s Z s Z s Z s Z s
(f +g)(t)dt = f (t) dt+ g(t) dt, (αf )(t) = α f (t) dt.
a a a a a
25
26 Linear Transformations
I(x) = x ∀ x ∈ V.
♦
What is a Linear Transformation? 27
♦
EXAMPLE 2.5 For each j ∈ {1, . . . , n}, the function fj : Fn → F
defined by fj (x) = xj for x = (α1 , . . . , αn ) ∈ Fn , is a linear func-
tional. ♦
More generally, we have the following example.
EXAMPLE 2.6 Let V be an n-dimensional space and let E =
{u1 , . . . , un } be a basis of V . For x = nj=1 αj uj ∈ V , and for each
P
j ∈ {1, . . . , n}, define fj : V → F by
fj (x) = αj .
(T1 + T2 )(x) = T1 x + T2 x,
(αT )(x) = αT x
for all x ∈ V . Then it is seen that L(V1 , V2 ) is a vector space with
its zero element as the zero operator O : V1 → V2 defined by
Ox = 0 ∀ x ∈ V1
n
X
αj = αi fi (uj ) = 0 ∀ j = 1, . . . , n.
i=1
m
X
T (uj ) = aij vi
i=1
n
X n
X m
X Xm X
n
T (x) = αj T uj = αj aij vi = aij αj vi . (∗)
j=1 j=1 i=1 i=1 j=1
m
X
Tx = (A~x)i vi .
i=1
Definition 2.6 The matrix [T ]E1 ,E2 is called the matrix representa-
tion of T , with respect to {E1 , E2 }. ♦
Clearly, the above discussion also shows that for every m × n
matrix A = (aij ), there exists a unique linear transformation T ∈
L(V1 , V2 ) such that [T ] = (aij ). Thus, there is a one-one correspon-
dence between L(V1 , V2 ) onto Fm×n , namely,
T 7→ [T ].
34 Linear Transformations
J2 T J1−1 x = [T ]x, x ∈ Fn ,
J2−1 [T ]J1 x = T x, x ∈ V1 .
Exercise 2.4 Prove the last statement. ♦
Exercise 2.5 Let V be an n-dimensional vector space and {u1 , . . . , un }
be an ordered basis of V . Let f be a linear functional on V . Prove
the following:
(i) There exists a unique (β1 , . . . , βn ) ∈ Fn such that
f (α1 u1 + . . . + αn un ) = α1 β1 + . . . + αn βn .
[T1 + T2 ]E1 ,E2 = [T1 ]E1 ,E2 + [T2 ]E1 ,E2 , [αT ]E1 ,E2 = α[T ]E1 ,E2 .
{Tij : i = 1 . . . , m; j = 1, . . . , n}
T (x1 , x2 , x3 ) = (x2 + x3 , x3 + x1 , x1 + x2 ).
R(T ) = {T x : x ∈ V1 }, N (T ) = {x ∈ V1 : T x = 0}
Hence,
r
X
T x− α j wj = 0
i=1
Pr
so that x − i=1 αj wj ∈ N (T ). Since E0 is a basis of N (T ), there
exist scalars β1 , . . . , βk such that
r
X k
X
x− α j wj = βj uj .
i=1 i=1
p(T ) = a1 I + a1 T + · · · + an T n .
so that
Thus, Te is linear.
T M = IY , M T = IX ,
(M T )−1 = M −1 T −1 .
T x = λ x,
T x = λx ⇐⇒ A~x = λ~x.
T (α1 , α2 , α3 ) = (α1 , α1 + α2 , α1 + α2 + α3 ).
T (α1 , α2 ) = (α1 + α2 , α2 ).
(T x)(t) = tx(t), x ∈ P.
Then Eig(T ) = ∅.
(vi) Let X be P[a, b] and A : V → X be defined by
d
(T x)(t) = x(t), x ∈ P.
dt
Then Eig(T ) = {0} and N (T ) = span {x0 }, where x0 (t) = 1 for all
t ∈ [a, b]. ♦
gλ + `λ − 1 ≤ m` .
a0 x + a1 T x + · · · + an T n x = 0.
p(t) = a0 + a1 t + · · · + ak tk , p(T ) = a0 I + a1 T + · · · + ak T k ,
we have
p(T )(x) = 0.
By fundamental theorem of algebra, there exist λ1 , . . . , λk in C such
that
p(t) = ak (t − λ1 )(t − λ2 ) . . . (t − λk ).
1
M.T. Nair: Multiplicities of an eigenvalue: Some observations, Resonance,
Vol. 7 (2002) 31-41.
44 Linear Transformations
Thus, we have
(T − λ1 I)(T − λ2 I) . . . (T − λk I)(x) = p(T )(x) = 0.
Hence, at least one of T − λ1 I, . . . , T − λk I is not one-one so that
at least one of λ1 , . . . , λk is an eigenvalue of A.
Theorem 2.13 Let λ1 , . . . , λk be distinct eigenvalues of a linear op-
erator T : V → V with corresponding eigenvectors u1 , . . . , uk , respec-
tively. Then {u1 , . . . , uk } is a linearly independent set.
Proof. We prove this result by induction. Clearly {u1 } is linearly
independent. Now, assume that {u1 , . . . , um } is linearly independent,
where m < k. We show that {u1 , . . . , um+1 } is linearly independent.
So, let α1 , . . . , αm+1 be scalars such that
α1 u1 + · · · + αm um + αm+1 um+1 = 0. (∗)
Applying T and using the fact that T uj = λj uj , we have
α1 λ1 u1 + · · · + αm λk um + αm+1 λm+1 um+1 = 0.
From (∗), multiplying by λm+1 , we have
α1 λm+1 u1 + · · · + αm λm+1 um + αm+1 λm+1 um+1 = 0.
Thus,
α1 (λ1 − λm+1 )u1 + · · · + αm (λm+1 − λm )um = 0.
Now, using the fact that {u1 , . . . , um } is linearly independent in V ,
and λ1 , . . . , λm , λm+1 are distinct, we obtain αj = 0 for j = 1, . . . , m.
Therefore, from (∗), αm+1 = 0. This completes the proof.
By the above theorem we can immediately infer that if V is finite
dimensional, then the eigenspectrum of every linear operator on V
is a finite set.
Definition 2.15 Let T ∈ L(V ) and V0 is a subspace of V . Then V0
is is said to be invariant under T if T (V0 ) ⊆ V0 , that is,
x ∈ V0 =⇒ T x ∈ V0 .
If V0 is invariant under T , then the restriction of T to the space V0
is the operator T0 ∈ L(V0 ) defined by
T0 x = T x ∀ x ∈ V0 .
♦
Eigenvalues and Eigenvectors 45
2.7.3 Diagonalizability
Definition 2.16 Suppose V is a finite dimensional vector space and
T : V → V is a linear operator. Then T is said to be diagonalizable
if V has a basis E such that [T ]E,E is a diagonal matrix. ♦
The proof of the following theorem is immediate (Write details!).
T (x1 , x2 , x3 ) = (x1 + x2 + x3 , x1 + x2 − x3 , x1 − x2 + x3 ).
T e1 = 0, T e2 = e1 , T e3 = e2 .
46 Linear Transformations
T e1 = e2 , T e2 = e3 , T e3 = 0.
T e1 = e3 , T e2 = e2 , T e3 = e1 .
♦
3
Inner Product Spaces
3.1 Motivation
In Chapter 1 we defined a vector space as an abstraction of the
familiar Euclidian space. In doing so, we took into account only two
aspects of the set of vectors in a plane, namely, the vector addition
and scalar multiplication. Now, we consider the third aspect, namely
the angle between vectors.
Recall from plane geometry that if ~x = (x1 , x2 ) and ~y = (y1 , y2 )
are two non-zero vectors in the plane R2 , then the angle θx,y between
~x and ~y is given by
x 1 y1 + x 2 y2
cos θx,y := ,
|~x| |~y |
h~x, ~y i = x1 y1 + x2 y2 .
x 7→ h~x, ~y i, ~x ∈ R2 ,
47
48 Inner Product Spaces
n
X
hx, yiE := αi β̄i .
i=1
n+1
X
hp, qi := p(τi )q(τi ), p, q ∈ Pn ,
i=1
hT x, T yi = hx, yi ⇐⇒ kT xk = kxk.
3.5 Orthogonality
Recall that the angle θx,y between vectors ~x and ~y in R2 is given by
h~x, ~y i2
cos θx,y := .
|~x| |~y |
• for x, y in V , x ⊥ y ⇐⇒ y ⊥ x, and
• 0 ⊥ x for all x ∈ V .
h~x, ~y i2
p~x,y := ~y
k~y k2
and it has less length atmost k~xk, that is, k~ px,y k ≤ k~xk. Thus, we
have
|h~x, ~y i2 | ≤ k~xk k~y k.
Orthogonality 53
Further, the vectors p~x,y and ~qx,y := ~x − p~x,y are orthogonal, so that
by Pythagoras theorem,
k~xk2 = k~
px,y k2 + k~qx,y k2 .
Exercise 3.8 (i) If the scalar field is R, then show that the converse
of the Pythagoras theorem holds, that is, if kx + yk2 = kxk2 + kyk2 ,
then x ⊥ y.
(ii) If the scalar field is C, then show that the converse of Pythago-
ras theorem need not be true .
[Hint: Take V = C with standard inner product, and for nonzero
real numbers α, β ∈ R, take x = α, y = i β.] ♦
Theorem 3.5 (Cauchy-Schwarz inequality) Let V be an inner
product space, and x, y ∈ V . Then
kx + yk ≤ kxk + kyk.
kx + yk2 = hx + y, x + yi
= hx, xi + hx, yi + hy, xi + hy, yi
= kxk2 + kyk2 + 2 Re hx, yi
≤ kxk2 + kyk2 + 2 |hx, yi|
≤ kxk2 + kyk2 + 2 kxk kyk
= (kxk + kyk)2 .
|hx, yi|
≤ 1.
kxkkyk
This relation motivates us to define the angle between any two nonzero
vectors x and y in V as
−1 |hx, yi|
θx,y := cos .
kxk kyk
S ⊥ := {x ∈ V : x ⊥ S}.
Since Z 2π Z 2π
cos(kt) dt = 0 = sin(kt) dt ∀ k ∈ Z,
0 0
we have for n 6= m,
and
n
X n
X n X
X n
2
kxk = hx, xi = αi ui , αj uj = αi ᾱj hui , uj i
i=1 j=1 i=1 j=1
n
X Xn
= |αi |2 = |hx, ui i|2 .
i=1 i=1
n
X
Proof. Let x ∈ V , and let y = hx, ui iui . By Theorem 3.10,
i=1
n
X
kyk2 = |hy, ui i|2 .
i=1
Note that hy, ui i = hx, ui i for all i ∈ {1, . . . , n}, i.e., hx−y, ui i = 0 for
all i ∈ {1, . . . , n}. Hence, hx − y, yi = 0. Therefore, by Pythagoras
theorem,
n
X
2 2 2 2
kxk = kyk + kx − yk ≥ kyk = |hx, ui i|2 .
i=1
ei nt
un (t) = √ , t ∈ [0, 2π].
2π
Then it is seen that
Z 2π
i (n−m)t 1 if n = m,
hun , um i = e dt =
0 0 6 m.
if n =
♦
Gram-Schmidt Orthogonalization 59
u2 = αu1 + v2 ,
which satisfy
k
X vk+1
vk+1 := uk+1 − huk+1 , wi iwi and wk+1 = .
kvk+1 k
i=1
V = V0 + W and V0 ⊥ W.
3.6.1 Examples
EXAMPLE 3.7 Let V = F3 with standard inner product. Consider
the vectors u1 = (1, 0, 0), u2 = (1, 1, 0), u3 = (1, 1, 1). Clearly,
u1 , u2 , u3 are linearly independent in F3 . Let us orthogonalize these
vectors according to the Gram-Schmidth orthogonalizaion procedure:
Take v1 = u1 , and
hu2 , v1 i
v2 = u2 − v1 .
hv1 , v1 i
Note that hv1 , v1 i = 1 and hu2 , v1 i = 1. Hence, v2 = u2 − v1 =
(0, 1, 0). Next, let
hu3 , v1 i hu3 , v2 i
v3 = u3 − v1 − v2 .
hv1 , v1 i hv2 , v2 i
Note that hv2 , v2 i = 1, hu3 , v1 i = 1 and hu3 , v2 i = 1 Hence, v3 =
u2 − v1 − v2 = (0, 0, 1). Thus,
Thus,
{(1, 1, 0), (−1/2, 1/2, 1), (−2/3, 2/3, −2/3)}
Z 1 Z 1
hu2 , v1 i = u2 (t)v1 (t) dt = t dt = 0.
−1 −1
Hence, we have v2 (t) = u2 (t) = t for all t ∈ [−1, 1]. Next, let
hu3 , v1 i hu3 , v2 i
v3 = u3 − v1 − v2 .
hv1 , v1 i hv2 , v2 i
Here,
Z 1 Z 1
2
hu3 , v1 i = u3 (t)v1 (t) dt = t2 dt = ,
−1 −1 3
Z 1 Z 1
hu3 , v2 i = u3 (t)v2 (t) dt = t3 dt = 0.
−1 −1
1
Hence, we have v3 (t) = t2 − 3 for all t ∈ [−1, 1]. Thus,
1
1, t, t2 −
3
hT x, yi = hAx, yi ∀ (x, y) ∈ V1 × V2 .
3.7 Diagonalization
In this section we show the diagonalizability of certain type operators
as promised at the end of Section 2.7.3.
Diagonalization 65
hT x, yi = hx, T yi ∀ x, y ∈ V.
♦
Suppose V is finite dimensional, E := {u1 , . . . , un } is an (ordered)
orthonormal basis of V , and T : V → V is a self-adjoint linear
transformation. Then we have the following:
hT x, xi ∈ R ∀ x ∈ V.
Hence, hx, yi = 0.
Therefore,
Au = λ u, Av = λ v.
Diagonalization 67
hT x, yi = hx, T yi = 0.
V0 = N (A − λ1 I) + . . . + N (A − λk I).
λ1 , . . . , λ1 , λ2 , . . . , λ2 , . . . , λk , . . . , λk
♦
We note the following (Verify!):
kf − pk ≤ kf − qk ∀ q ∈ Pn .
Here, k.k is a norm on C[a, b]. Now the question is whether such
a polynomial exists, and if exists, then is it unique; and if there is
a unique such polynomial, then how can we find it. These are the
issues that we discuss in this section, in an abstract frame work of
inner product spaces.
Definition 3.12 Let V be an inner product spaceand V0 be a sub-
space of V . Let x ∈ V . A vector x0 ∈ V0 is a called a best approx-
imation of x from V0 if
kx − x0 k ≤ kx − vk ∀ v ∈ V0 .
Hence
kx − x0 k ≤ kx − vk ∀ v ∈ V0 ,
70 Inner Product Spaces
Pn
Proof. Clearly, x0 := i=1 hx, ui iui satisfies the hypothesis of
Proposition 3.24.
Z 1
hx − x0 , u2 i = (t3 − a0 t − a1 t2 )dt = 0.
0
That is
Z 1
(t2 − a0 − a1 t)dt = [t3 /3 − a0 t − a1 t2 /2]10 = 1/3 − a0 − a1 /2 = 0,
0
Z 1
(t3 −a0 t−a1 t2 )dt = [t4 /4−a0 t2 /2−a1 t3 /3]10 = 1/4−a0 /2−a1 /3 = 0.
0
Exercise
R1 3.17 Let V = C[0, 1] over R with inner product hx, ui =
0 x(t)u(t)dt. Let V0 = P3 . Find best approximation for x from V0 ,
where x(t) is given by
(i) et , (ii) sin t, (iii) cos t, (iv) t4 . ♦
kAx0 − yk ≤ kAu − yk ∀ u ∈ V1 .
♦
It is obvious that x0 ∈ V1 is a best approximate solution of Ax =
y if and only if y0 := Ax0 is a best approximation of y from the
range space R(A). Thus, from Proposition 3.24, we can conclude the
following.
AT Ax0 = AT y.
RT QT QRx0 = RT QT y.
Rx0 = QT y.
Appendix 75
Rx = QT y.
3.11 Appendix
Another proof for Theorem 3.20. Assuming F = R, another
way of proving Theorem 3.20 is as follows: Consider a new vector
space Ve := {u + iv : u, v ∈ V } over C. For u + iv, u1 + iv1 , u2 + iv2
in Ve and α + iβ) ∈ C with (α, β) ∈ R2 the addition and scalar
multiplication are defined as
Define Te : Ve → Ve by
Te(u + iv) := T u + iT v.
76
Norms of Vectors and Matrices 77
Exercise 4.2 Show that there exists no constant c > 0 such that
kxk∞ ≤ c kxk1 for all x ∈ C[a, b].
defines a norm on the space Rn×n . Since this norm is associated with
the norm of the space Rn , and since a matrix can be considered as
a linear operator on Rn , the above norm on Rn×n is called a matrix
norm associated with a vector norm.
The above norm has certain important properties that other norms
may not have. For example, it can be seen that
• kAxk ≤ kAk kxk ∀x ∈ Rn ,
• kAxk ≤ c kxk ∀x ∈ Rn =⇒ kAk ≤ c,.
Moreover, if A, B ∈ Rn×n and if I is the identity matrix, then
• kABk ≤ kAk kBk, kIk = 1.
Thus, kAk1 ≤ max1≤j≤n ni=1 |aij |.PAlso, note that kAej k1 = ni=1 |aij |
P P
n
Pn i=1 |aij | ≤ kAk1 for every j ∈
for every j ∈ {1, . . . , n} so that
{1, . . . , n}. Hence, max1≤j≤n i=1 |aij | ≤ kAk1 . Thus, we have
shown that
X n
kAk1 = max |aij |.
1≤j≤n
i=1
Since
Xn X n n
X
aij xj ≤ |aij | |xj | ≤ kxk∞ |aij |,
j=1 j=1 j=1
it follows that
n
X
kAxk∞ ≤ max |aij | kxk∞ .
1‘i≤n
j=1
Norms of Vectors and Matrices 79
n
X
From this we have kAk∞ ≤ max |aij |. Now, let i0 ∈ {1, . . . , n}
1‘i≤n
j=1
n
X n
X
be such that max |aij | = |ai0 j |, and let x0 = (α1 , . . . , αn ) be
1≤i≤n
j=1 j=1
|ai0 j |/ai0 j 6 0,
if ai0 j =
such that αj = Then kx0 k∞ = 1 and
0 if ai0 j =6 0.
n
X Xn
|ai0 j | = ai0 j αj = |(Ax0 )io | ≤ kAx0 k∞ ≤ kAk∞ .
j=1 j=1
n
X n
X
Thus, max |aij | = |ai0 j | ≤ kAk∞ . Thus we have proved that
1≤i≤n
j=1 j=1
n
X
kAk∞ = max |aij |.
1≤i≤n
j=1
n X
X n 1/2
kAk2 ≤ |aij |2 .
i=1 j=1
1 2 3
Exercise 4.4 Find kAk1 , kAk∞ , for the matrix A = 2 3 4 .
3 2 1
Ax̃ = b̃.
kAxk kb − b̃k
kx− x̃k ≤ kA−1 kb− b̃k = kA−1 kb− b̃k ≤ kAk kA−1 k kxk,
kbk kbk
kA−1 bk kx − x̃k
kb−b̃k ≤ kAk kx−x̃k = kAk kx−x̃k ≤ kAk kA−1 k kbk.
kxk kxk
Thus, denoting the quantity kAk kA−1 k by κ(A),
kx − x̃k kb − b̃k
= κ(A) ,
kxk kbk
where x, x̃ are such that Ax = b and Ax̃ = b̃. To see this, let x0 and
u be vectors such that
and let
b := Ax0 , b̃ := b + u, x̃ := x0 + A−1 u.
82 Error Bounds and Stability of Linear Systems
kx − x̃k kA − Bk kb − b̃k
≤ 2κ(A) + .
kxk kAk kbk
84
85
xn = f (xn−1 ), n = 1, 2, . . . .
for some constant ρ satisfying 0 < ρ < 1. Then f has a unique fixed
point. In fact, for any x0 ∈ S, if we define
xn = f (xn−1 ), n = 1, 2, . . . ,
ρm
kxn − xm k ≤ kx1 − x0 k ∀n > m,
1−ρ
ρn
kx − xn k ≤ kx1 − x0 k ∀n ∈ N.
1−ρ
Proof. Let x0 ∈ S, and define
xn = f (xn−1 ), n = 1, 2, . . . .
Then
so that
kxn+1 − xn k ≤ ρn kx1 − x0 k ∀n ∈ N.
87
Thus, under the additional assumptions (i) and (ii), we can generate
the iterations with any x0 ∈ Dr .
Remark 5.2 In order to have certain accuracy of the approximation,
say for the error kx − xn k to at most ε > 0, we have to take n large
enough so that
ρn
kx1 − x0 k < ε,
1−ρ
that is, error kx − xn k ≤ ε > 0 for all n satisfying
log kx1 − x0 k/ε(1 − ρ)
n≥ .
log(1/ρ)
x(k) = Cx(k−1) + d, n = 1, 2, . . . .
F (x) = Cx + d, x ∈ Rn .
Then we have
Ax = b ⇐⇒ x = (I − A)x + b.
x(k) = (I − A)x(k−) + b, n = 1, 2, . . .
x = A−1 −1
1 b − A1 A2 x.
A1 x(k) = b − A2 x(k−1) , k = 1, 2, . . . .
(k) 1 h X (k−1)
i
xi = bi − aij xj , k = 1, 2, . . .
aii
j6=i
We see that
1 1 1
9 0 0 0 9 9
A−1
1 =
0 1
10 0 so that A−1
1 A2 =
1
5 0 3
10
.
1 3 4
0 0 11 11 11 0
ρk 99 47 k (1)
kx − x(k) k1 ≤ kx(1) − x(0) k = kx − x(0) k.
1−ρ 52 99
5.1.2 Gauss-Siedel Method
Let A = (aij ) be an n × n matrix. In this method also, we assume
that aii 6= 0 for all i = 1, . . . , n. In this case we view the system
Ax = b as
A1 x = b − A2 x,
Iterative Methods for Solving Ax = b 91
(k) 1 h X (k)
X (k−1)
i
xi = bi − aij xj − aij xj , k = 2, . . .
aii
j<i j>i
We see that
1 1 1
9 0 0 0 9 9
A−1
1 =
−1
45
1
10 0 so that −1 1
A1 A2 = 0 − 45 5
− 18 .
1 2 1 1 13
− 45 − 55 11 0 − 45 − 99
f (x)f 00 (x)
g 0 (x) := , x ∈ [a, b].
[f 0 (x)]2
Now, by mean value theorem, for every x, y ∈ [a, b], there exists ξx,y
in the interval whose end points are x and y, such that
ρn
|x∗ − xn | ≤ ρ|x∗ − xn−1 | ≤ |x1 − x0 | ∀ n ∈ N.
1−ρ
f 00 (ξn )
0 = f (x∗ ) = f (xn ) + (x∗ − xn )f 0 (xn ) + (x∗ − xn )2
2
94 Fixed Point Iterations for Solving Equations
so that 00
f (xn ) ∗ ∗ 2 f (ξn )
0= + (x − x n ) + (x − x n )
f 0 (xn ) 2f 0 (xn )
Now, by the definition of xn+1 ,
f 00 (ξn )
0 = (xn − xn+1 + (x∗ − xn ) + (x∗ − xn )2
2f 0 (xn )
so that
f 00 (ξn )
x∗ − xn+1 = (x∗ − xn )2 .
2f 0 (xn )
From the above relation, it is clear that if J0 is as in Theorem
5.4, and if we know that there exists a constant κ > 0 such that
|f 00 (x)/2f 0 (y)| ≤ κ for all x, y ∈ J0 , then
6.1 Interpolation
The idea of interpolation is to find a function ϕ which takes certain
prescribed values β1 , β2 , . . . , βn at a given set of points t1 , t2 , . . . , tn .
In application the values β1 , β2 , . . . , βn may be values of certain un-
known function f at t1 , t2 , . . . , tn respectively. The function ϕ is to
be of some simple form for computational purposes. Thus, the in-
terpolation problem is to find a function ϕ such that ϕ(ti ) = βi ,
i = 1, . . . , n.
Usually, one looks for ϕ in the span of certain known functions
u1 , . . . , un . Thus, the interpolation P problem is to find scalars α1 , . . . , αn
such that the function ϕ := nj=1 αj uj satisfies ϕ(ti ) = βi for i =
1, . . . , n, i.e., to find α1 , . . . , αn such that
n
X
αj uj (ti ) = βi , i = 1, . . . , n.
j=1
Obviously, the above problem has a unique solution if and only if the
matrix [uj (ti )] is invertible. Thus we have the following theorem.
96
Interpolation 97
if tn ≤ t ≤ b,
1
t−tn−1
un (t) = if tn−1 ≤ t ≤ tn ,
tn −tn−1
0, elsewhere,
and for 2 ≤ j ≤ n − 1,
t−tj−1
tj −tj−1 if tj−1 ≤ t ≤ tj ,
t−tj+1
uj (t) = tj −tj+1 if tj ≤ t ≤ tj+1 ,
0, elsewhere,
98 Interpolation and Numerical Integration
Z 1
Error = ln(2) − ϕ(t)dt ' −0.01518
0
b
(t − c)(t − b)
Z
h
dt = ,
a (a − c)(a − b) 3
b
(t − a)(t − b)
Z
4h
dt = ,
a (c − a)(c − b) 3
b
(t − a)(t − c)
Z
h
dt = .
a (b − a)(b − c) 3
Hence,
Z b
hh i
ϕ(t)dt = f (a) + 4f (c) + f (b) .
a 3
This quadrature rule is called the Simpson’s rule.
EXAMPLE 6.4 Consider f (t) = 1/(1 + t) for 0 ≤ t ≤ 1. Taking
n = 2, hn = 1/2 and
Z 1
hh i
ϕ(t)dt = f (a) + 4f (c) + f (b)
0 3
1h i 1h1 2 1i
= f (0) + 4f (1/2) + f (1) = +4 +
6 6 2 3 2
25
= ' 0.69444.
36
Z 1
Error = ln(2) − ϕ(t)dt ' −0.001293
0
Numerical Integration 103
Thus,
Z b
hn h
ϕ(t)dt = f (a0 ) + 4f (a1 ) + 2f (a2 ) + 4f (a3 ) + 2f (a4 ) +
a 3
i
. . . + 2f (an−2 ) + 4f (an−1 ) + f (an )
k k−1
hn h X X i
= f (a0 ) + 4f (a2i−1 ) + 2f (a2i ) + f (an )
3
i=1 i=1
104
105
10. Let V be the space of all sequences of real numbers, and let
`1 (N) be the set of all absolutely convergent real sequences.
Show that `1 (N) is a subspace of V .
11. Let V be the space of all sequences of real numbers, and let
`∞ (N) be the set of all bounded sequences of real numbers.
Show that `∞ (N) is a subspace of the space of V .
14. Give an example to show that union of two subspaces need not
be a subspace.
16. Let V be a vector space. Show that the the following hold.
(i) Let S be a subset of V . Then span S is the intersection of
all subspaces of V containing S.
18. For each λ in the open interval (0, 1), let uλ = (1, λ, λ2 , . . .).
Show that uλ ∈ `1 for each ∈ (0, 1), and {uλ : 0 < λ < 1} is a
linearly independent subset of `1 .
20. Let e1 = (1, 0, 0), e2 = (0, 1, 0), e3 = (0, 0, 1). What is the span
of {e1 + e2 , e2 + e3 , e3 + e1 }?
23. Let V be a vector space. Show that the the following hold.
29. Show that vectors u = (a, c), v = (b, d) are linearly independent
in R2 iff ad − bc 6= 0. Can you think of a generalization to n
vectors in Rn .
39. For each k ∈ N, let Fk denotes the set of all column k-vectors,
i.e., the set of all k × 1 matrices. Let A be an m × n matrix of
scalars with columns a1 , a2 , . . . , an . Show the following:
46. Check whether the functions T in the following are linear trans-
formations:
[A+B]E1 ,E2 = [A]E1 ,E2 +[B]E1 ,E2 , [αA]E1 ,E2 = α[A]E1 ,E2 .
(a) E1 = {(1, 0, 0), (0, 1, 0), (0, 0, 1)}, E2 = {(1, 0, 0), (1, 1, 0), (1, 1, 1)}
(b) E1 = {(1, 0, 0), (1, 1, 0), (1, 1, 1)}, E2 = {(1, 0, 0), (0, 1, 0), (0, 0, 1)}
(c) E1 = {(1, 1, −1), (−1, 1, 1), (1, −1, 1)},
E2 = {(−1, 1, 1), (1, −1, 1), (1, 1, −1)
(a) E1 = {1, t, t2 , t3 }, E2 = {1 + t, 1 − t, t2 }
(b) E1 = {1, 1 + t, 1 + t + t2 , t3 }, E2 = {1, 1 + t, 1 + t + t2 }
(c) E1 = {1, 1 + t, 1 + t + t2 , 1 + t + t2 + t3 }, E2 = {t2 , t, 1}
(a) E1 = {1 + t, 1 − t, t2 }, E2 = {1, t, t2 , t3 },
(b) E1 = {1, 1 + t, 1 + t + t2 }, E2 = {1, 1 + t, 1 + t + t2 , t3 },
(c) E1 = {t2 , t, 1}, E2 = {1, 1 + t, 1 + t + t2 , 1 + t + t2 + t3 },
65. Check whether the following are inner product on the given
vector spaces:
68. Let h·, ·i1 and h·, ·i2 are inner products on a vector space V .
Show that
S ⊥ := {x ∈ V : hx, ui = 0 ∀u ∈ S}.
Show that
(a) S ⊥ is a subspace of V .
(b) V ⊥ = {0}, {0}⊥ = V .
(c) S ⊂ S ⊥⊥ .
(d) If V is finite dimensional and V0 is a subspace of V , then
V0⊥⊥ = V0 .
75. Find the best approximate solution (least square solution) for
the system Ax = y in each of the following:
3 1 1
(a) A = 1 2 ; y = 0 .
2 −1 −2
1 1 1 0
−1 0 1 1
(b) A =
1 −1
; y=
−1 .
0
0 1 −1 −2
116 Additional Exercises
1 1 3 1
−1 0 5
−1
(c) A =
0 1 −2
; y=
3 .
1 −1 1 −2
1 0 1 0
for all x ∈ Rk .
(c) Compute kxk∞ , kxk2 , kxk1 for x = (1, 1, 1) ∈ R3 .
kA−1 k
kB −1 k ≤ .
1 − kA − Bk kA−1 k
117
89. Let t1 , . . . , tn be distinct points in [a, b]. Show that for every
x ∈ C[a, b], there exists a unique polynomial p(t) of degree
atmost n − 1 such that p(tj ) = x(tj ) for j = 1, . . . , n.