0% found this document useful (0 votes)
174 views59 pages

Revision Notes - MA2101

The document is a set of revision notes covering topics in linear algebra, including: 1. Definitions of fields, vector spaces, subspaces, and linear combinations. 2. Properties of fields including uniqueness of identities and inverses. 3. Theorems regarding subspaces including their intersections being subspaces and sums of subspaces being the smallest subspace containing both. 4. Definitions of linear spans, sums of subspaces, and their properties including spans being the smallest subspace containing a set and sums being subspaces.

Uploaded by

johndoeunspec
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
174 views59 pages

Revision Notes - MA2101

The document is a set of revision notes covering topics in linear algebra, including: 1. Definitions of fields, vector spaces, subspaces, and linear combinations. 2. Properties of fields including uniqueness of identities and inverses. 3. Theorems regarding subspaces including their intersections being subspaces and sums of subspaces being the smallest subspace containing both. 4. Definitions of linear spans, sums of subspaces, and their properties including spans being the smallest subspace containing a set and sums being subspaces.

Uploaded by

johndoeunspec
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 59

Revision notes - MA2101

Ma Hongqiang
April 27, 2017

Contents
1 Vector Spaces over a Field 2

2 Vector Subspaces 6

3 Linear Spans and Direct Sums of Subspaces 7

4 Linear Independence, Basis and Dimension 10

5 Row Space and Column Space 14

6 Quotient Spaces and Linear Transformations 18

7 Representation Matrices of Linear Transformations 24

8 Eigenvalue and Cayley-Hamilton Theorem 29

9 Minimal Polynomial and Jordan Canonical Form 37

10 Quadratic Forms, Inner Product Spaces and Conics 48

11 Problems 59

1
1 Vector Spaces over a Field
Definition 1.1 (Field, Rings, Groups).
Let F be a set containing at least two elements and equipped with the following two binary
operations +(the addition, or plus) and ×(the multiplication, or times), where F × F :=
{(x, y) | x, y ∈ F} is the product set of F with iteself:

+ :F × F → F
(x, y) 7→ x + y;
× :F × F → F
(x, y) 7→ x × y;

Axiom(0) Addition and multiplication are well defined on F in the sense that:

∀x ∈ F, ∀y ∈ F ⇒ x + y ∈ F
∀x ∈ F, ∀y ∈ F ⇒ xy ∈ F

Namely, the operation map +(resp. ×) takes element (x, y) in the domain F × F to some ele-
ment x + y(resp. xy) in the codomain F. The quintuple (F, +, 0; ×, 1) with two distinguished
elements 0 (the additive identity) and 1 (the multiplicative identity), is called a field if the
following Eight Axioms (and also Axiom (0)) are satisfied.
(1) Existence of an additive identity 0F or simply 0:

x + 0 = x = 0 + x, ∀x ∈ F

(2) (Additive) Associativity:

(x + y) + z = x + (y + z), ∀x, y, z ∈ F

(3) Additive Inverse


for every x ∈ F, there is an additive inverse −x ∈ F of x such that

x + (−x) = 0 = (−x) + x

(4) Existence of a multiplicative identity 1F or simply 1:

x1 = x = 1x, ∀x ∈ F

(5) (Multiplicative) Associativity:

(xy)z = x(yz), ∀x, y, z ∈ F

(6) Multiplicative Inverse for nonzero element:


for every 0 6= x ∈ F, there is a multiplicative inverse x−1 ∈ F such that

xx−1 = 1 = x−1 x.

2
(7) Distributive Laww:
(x + y)z = xz + yz, ∀x, y, z ∈ F
z(x + y) = zx + zy, ∀x, y, z ∈ F

(8) Commutativity for addition and multiplication:

x + y = y + x, ∀x, y ∈ F
xy = yx, ∀x, y ∈ F

The triplet (F, +, ×) with only the Axioms (1)—(5) and (7)—(8) satisfied is called a (com-
mutative) ring.
The pair (F, +) with only Axioms (1)—(3) satisfied by its binary operation +, is called an
(additive) group.

Notation 1.1 (about F× ). For a field (F, +, 0; ×, 1), we use F× to denote the set of nonzero
elements in F:
F× := F \ {0}

Definition 1.2 (Polynomial ring).


Let (F, +, 0; ×, 1) be a field (or a ring), e.g. F = Z, Q, R, C.
n
X
g(x) = ai xi = an xn + an−1 xn−1 + · · · + a1 x + a0
i=0

with the leading coefficient an = 6 0, is called a polynomial of degree n ≥ 0, in one


variable x and with coefficients ai ∈ F. Let
d
X
F[x] := { bj xj | d ≥ 0, bj ∈ F}
j=0

be the set of all polynomials in one variable x and with coefficients in F.

Theorem 1.1 (Uniqueness of identity and inverse). Let F be a field.

(1) F has only one additive identity 0.

(2) F has only one multiplicative identity 1.

(3) Every x ∈ F has only one additive inverse −x.

(4) Every x ∈ F× has only one multiplicative inverse x−1 .

Theorem 1.2 (Properties of F).

(1) (Cancelation Law) Let b, x, y ∈ F. Then

b+x=b+y ⇒x=y

3
(2) (Killing Power of 0)
0x = 0 = x0, ∀x ∈ F

(3) In F, we have
0F 6= 1F

(4) If x ∈ F× , then its multiplicative inverse

x−1 ∈ F×

(5) If x + x0 = 0, then x and x0 are multiplicative inverse to each other.

(6) If xx00 = 1 then x and x00 are multiplicative inverse to each other.

Definition 1.3 (Vector Space).


Let F be a field and V a non-empty set, with a binary vector addition operation

+:V ×V →V
(v1 , v2 ) 7→ v1 + v2

and scalar multiplication operation

×:V ×V →V
(c, v) 7→ cv

Axiom(0) These two operations are well defined in the sense that

∀vi ∈ V ⇒ v1 + v2 ∈ V
∀c ∈ F, ∀v ∈ V ⇒ cv ∈ V

(V, +) is a vector space over the field F if the following Seven Axioms are satisfied.

(1) Existence of zero vector 0V :

v + 0 = v = v + 0, ∀v ∈ V

(2) (Additive) Associativity:

(u + v) + w = u + (v + w), ∀u, v, w ∈ V

(3) Additive Inverse:


for every v ∈ V , there is an additive inverse −v of v such that

v + (−v) = 0 = (−v) + v

(4) The effect of 1 ∈ F on V :


1v = v, ∀v ∈ V

4
(5) (Multiplicative) Associativity:

(ab)v = a(bv), ∀a, b ∈ F, v ∈ V

(6) Distributive Law:


(a + b)v = av + bv, ∀a, b ∈ F, v ∈ V
a(u + v) = au + av, ∀a ∈ F, u, v ∈ V

(7) Commutativity for the vector addition:

u + v = v + u, ∀u, v ∈ V

Remark: Every vector space V contains the zero vector 0V .

5
2 Vector Subspaces
Definition 2.1 (Subspace).
Let V be a vector space over a field F. A non-empty subset W ⊆ V is called a vector
subspace of V if the following two conditions are satisfied:

(CA) (Closed under vector Addition)

∀wi ∈ W ⇒ w1 + w2 ∈ W

(CS) (Closed under Scalar Multiplication)

∀a ∈ F, ∀w ∈ W ⇒ aw ∈ W

Obvious subspace of V are V and {0}.


Remark: Every vector subspace W of V contains the zero vector 0W .

Definition 2.2 (Linear Combination).


Let V be a vector space over a field F and W ⊆ V a non-empty subset. Then the following
are equivalent:

(i) W is a vector subspace of V , i.e. W is closed under vector addition and scalar multi-
plication, in the sense of Definition of Subspace.

(ii) W is closed under linear combination:

∀ao ∈ F, ∀wi ∈ W ⇒ a1 w1 + a2 w2 ∈ W

(iii) W together with the vector addition + and the scalar multiplication ×, becomes a
vector space.

Theorem 2.1 (Intersection of Subspace being a subspace).


Let V be a vector space over a field F and let Wα ⊆ V (α ∈ I) be vector subspaces of V .
Then the intersection
∩α∈I Wα
is again a vector subspace of V

Remark: Union of subspaces may not be a subspace. However, union of subspaces is closed
under scalar multiplication.

6
3 Linear Spans and Direct Sums of Subspaces
Definition 3.1 (Linear combination and Linear Span).
Let V be a vector space over a field F. A vector v ∈ V is called a linear combination of
some vectors vi ∈ V (1 ≤ i ≤ s) if

v = a1 v1 + a2 v2 + · · · + as vs

for some scalars ai ∈ F. Let S ⊆ V be a non-empty subset. The subset Span(S) :=

{v ∈ V | v is a linear combination of some vectors in S}

of V is called the vector subspace of V spanned by the subset S.

Theorem 3.1 (Span being a subspace).

(i) The subset Span(S) of V is indeed a vector subspace of V .

(ii) Span(S) is the smallest vector subspace of V containing the set S:


firstly, Span(S) is a vector subspace of V containing S;
secondly, if W is another vector subspace of V containing S, then W ⊇ Span(S)

Definition 3.2 (Sum of subspaces).


Let V be a vector space over a field F, and let U and W be vector subspaces of V . The
subset
U + W = u + w | u ∈ U, w ∈ W
is called the sum of the subspaces U and W .

Theorem 3.2 (Sum being a subspace).


Let U and W be vector subspaces of a vector space V over a field F. For the sum U + W ,
we have:

1. U + W = Span(U ∪ W )

2. U + W is indeed a vector subspace of V .

3. U + W is the smallest vector subspace of V containing both U and W :


first, U + W is a vector subspace of V containing both U and W ; secondly, if T is
another vector subspace of V containing both U and W then, T ⊇ U + W .

Note 1: Let U and W be two vector subspaces of a vector space V over a field F. Then the
following are equivalent.

1. The union U ∪ W is a vector subspace of V .

2. Either U ⊆ W or W ⊆ U .

7
Definition 3.3 (Sum of many subspaces).
Let V be a vector space over a field F and let Wi (1 ≤ i < s) be vector subspaces of V . The
subset
= W1 + · · · + Ws
s s
X
X
Wi = { wi | wi ∈ W }
i=1 i=1
= {w1 + · · · + ws | wi ∈ Wi }

Theorem 3.3 (Sum of many being a subspace).


Let Wi (1 ≤ i < s) be vector subspaces of a vector space V over a field F. For the sum
P s
i=1 Wi , we have
Ps s
1. i=1 Wi = Span(∪i=1 Wi )
Ps
2. i=1 Wi is indeed a vector subspace of V .
Ps
3. i=1 Wi is the smallest vector subspace of V containing all Wi .

Definition 3.4 (Direct Sum of Subspaces).


Let V be a vector space over a field F and let W1 , W2 be vector subspace of V . We say that
the sum W1 + W2 is a direct sum of two vector subspaces W1 , W2 if the intersection

W1 ∩ W2 = {0}

In this case, we denote W1 + W2 as W1 ⊕ W2 .


We write W = W1 ⊕ W2 if W is a direct sum of W1 and W2 .

Theorem 3.4 (Equivalent Direct Sum Definition).


Let W1 and W2 be two vector subspaces of a vector space V over a field F. Set W := W1 +W2 .
Then the following are equivalent.

1. We have
W1 + W2 = W1 ⊕ W2
i.e., W1 + W2 is a direct sum of W1 , W2 ,i.e., W1 ∩ W2 = {0}

2. (Unique expression condition) Every vector w ∈ W can be expressed as

w = w1 + w2

for some wi ∈ Wi and such expression of w is unique.

Definition 3.5 (Direct Sum of Many Subspaces).


P overa field F and let Wi (1 ≤ i ≤ s; s ≥ 2) be vector subspaces of V .
Let V be a vector space
We say that the sum si=1 Wi is a direct sum of vector subspaces Wi if the intersection
k−1
!
X
Wi ∩ Wk = {0} (2 ≤ ∀k ≤ s)
i=1

8
Theorem 3.5 (Equivalent Direct Multiple Sum Definition).
i (1 ≤ i ≤ s, s ≥ 2) be vector subspaces of a vector space V over a field F. Set
Let WP
W := si=1 Wi . Then the following are equivalent.

1. We have
W1 + · · · + Ws = W1 ⊕ · · · ⊕ Ws
Ps
i.e., i=1 Wi is a direct sum of Wi .

2. !
X
Wi ∩ Wl = {0} (∀1 ≤ l ≤ s)
i6=j

3. (Unique expression condition) Every vector w ∈ W can be expressed as

w = w1 + · · · + ws

for some wi ∈ Wi and such expression of w is unique.

9
4 Linear Independence, Basis and Dimension
Definition 4.1 (Linear (in)dependence).
Let V be a vector space over a field F. Let T be a (not necessarily finite) subset of V and
let
S = {v1 , . . . , vn }
be a finite subset of V .
(1) We call S a linear independent set or L.I., if the vector equation below

x1 v1 + · · · + xm vm = 0

has only the so called trivial solution

(x1 , . . . , xm ) = (0, . . . , 0)

(2) We call S a linear dependent set or L.D. if there are scalars a1 , . . . , am in F which
are not all zero (i.e. (a1 , . . . , am ) 6= (0, . . . , 0)) such that

a1 v1 + · · · + am vm = 0

(3) The set T is a linearly independent set if every non-empty finite subset of T is
linearly independent. The set T is a linearly dependent set if at least one non-
empty finite subset of T is linearly dependent.
Theorem 4.1 (L.D./L.I. Inheritance).
(1) Let S1 ⊆ S2 . If the smaller set S1 is linearly dependent then so is the larger set S2 .
Equivalently, if the larger set S2 is linearly independent then so is the smaller set S1 .

(2) {0} is a linearly dependent set.

(3) If 0 ∈ S, then S is a linearly dependent set.


Definition 4.2 (Equivalent L.I./L.D. Definitions).
Let S = {v1 , . . . , vm } be a finite subset of a vector space V over a field F. Then we have:
(1) Let |S| ≥ 2. Then S is a linear dependent set if and only if some vk ∈ S is a linear
combination of the others, i.e. there are scalars

a1 , . . . , ak−1 , ak+1 , . . . , am

in F (with all these scalars vanishing allowed) such that


X
vk = ai vi = a1 v1 + · · · + ak−1 vk−1 + ak+1 vk+1 + · · · + am vm
i6=k

(2) Let |S| ≥ 2. Then S is linearly independent if and only if no vk ∈ S is a linear


combination of others.

10
(3) Suppose that S = {v1 } (a single vector). Then S is linearly dependent if and only if
v1 = 0. Equivalently, S is linearly independent if and only if v1 6= 0.

(4) Suppose that S = {v1 , v2 } (two vectors). Then S is linearly dependent if and only if
one of v1 , v2 is a scalar multiple of the other. Equivalently, S is linearly independent
if and only if neither one of v1 , v2 is a scalar multiple of the other.
Definition 4.3 (Basis, (in)finite demension).
Let V be a nonzero vector space over a field F. A subset B of V is called a basis if the
following two conditions are satisfied.
(1) (Span) V is spanned by B: V = Span(B)

(2) (L.I.) B is a linearly independent set.


If V has a basis B with cardinality |B| < ∞ we say that V is finite dimensional and
define the dimension of V over the field F as the cardinality of B:

dimF V := |B|

Otherwise, V is called infinite-dimensional.


If V equals the zero vector space {0}, we define

dim{0} = 0

Theorem 4.2 (Equivalent Basis Definition I).


Let B = {v1 , . . . , vn }(with vi 6= 0V ) be a finite subset of a vector space V over a field F.
Then the following are equivalent.
(1) B is a basis of V .

(2) (Unique expression condition) Every vector v ∈ V can be expressed as

v = a1 v1 + · · · + an vn

for some scalars ai ∈ F and such expression of v is unique.

(3) V has the following direct sum decomposition:

= Span{v1 } ⊕ · · · ⊕ Span{vn }
V
= Fv1 ⊕ · · · ⊕ Fvn

Theorem 4.3 (Deriving a basis from a spanning set).


Suppose that a nonzero vector space V over a field F is spanned by a finite subset B =
{v1 , . . . , vs }, then we have:
(1) There is a subset B1 ⊆ B such that B1 is a basis of V . In particular

dimF V = B1 ≤ B

11
(2) Let B2 be a maximal linearly independent subset of B: first B2 is L.I. and secondly
every subsets B3 of B larger than B2 is L.D. Then B2 is a basis of V = Span(B).

Theorem 4.4 (Dimension being well Defined).


Let B = {v1 , . . . , vn } be a basis of a vector space V over a field F. Then we have:

(1) Suppose that S is a subset of V with |S| > n = |B|. Then S is L.D.

(2) Suppose that T is a subset of V with |T | < n. Then T does not span V .

(3) Suppose that B 0 is another basis of V . Then |B 0 | = |B|. So the dimension dimF V (=
|B|) of V depends only on V , but not on the choice of its basis.
In other words, dimF V is well defined.

Theorem 4.5 (Expanding an L.I. set).


Let B be a L.I. subset of a vector space V over a field. Then exactly one of the following
two cases is true.

(1) B spans V and hence B is a basis of V .

(2) Let w ∈ V \ Span(B)( and hence w ∈


/ B). Then

B ∪ {w}

is a L.I. subset of V .

In particular, if V is of finite dimension n, then one can find n − |B| vectors

w|B|+1 , · · · , wn

in V Span(B) such that a


B {w|B|+1 , · · · , wn }
is a basis of V .

Theorem 4.6 (Equivalent Basis Definition II).


Let B be a subset of vector space V of finite dimension dimm athbbF V = n ≥ 1. Then the
following are equivalent.

(1) B is a basis of V .

(2) B is L.I. and |B| = n.

(3) B spans V and |B| = n.

Theorem 4.7 (Basis of a direct sum).


Let V be a (not necessarily finite-dimensional) vector space over a field F.

12
(1) Suppose that B is a basis of V . Decompose it as a disjoint union
a a
B = B1 B2 · · · Bs

of non-empty sets Bi . Then Bi is a basis of Wi := Span(Bi ) and

V = W1 ⊕ · · · ⊕ Ws

is a direct sum of nonzero vector subspaces Wi of V .

(2) Conversely, suppose that


V = W1 ⊕ · · · ⊕ Ws
is a direct sum of nonzero vector subspaces Wi of V . Let Bi be a basis of Wi . Then
a a
B = B1 B2 · · · Bs

is a basis of V and a disjoint union of non-empty sets Bi .

(3) In particular, if
V = W1 ⊕ · · · ⊕ Ws
is a direct sum, then
s
X
dimm athbbF V = dimF Wi
i=1

13
5 Row Space and Column Space
Definition 5.1 (Column/Row Space, Nullspace, Nullity, Range of A).
Let  
a11 · · · a1n
A = (ai j) =  ... .. .. 

. . 
am1 · · · amn
be an m × n matrix with entries in a field F. Let
   
a11 a1n
Col(A) := Span{c1 :=  ...  , . . . , cn :=  ... }
   
am1 amn
be the column space of A, and let
R(A) := Span{r1 := (a11 , . . . , a1n ), . . . , rm := (am1 , . . . , amn )}
be the row space of A so that we can write
 
r1
A = (c1 , . . . , cm ) =  ... 
 
rm
The range of A is defined as
R(A) = {AX | X ∈ Fnc }
The nullity of A is defined as the dimension of the nullspace or kernel
Ker(A) := Null(A) := {X ∈ Fnc | AX = 0}
i.e.
nullity(A) = dim Null(A)
Theorem 5.1 (Rank of Matrix, Matrix Dimension Theorem).

(1) The range equals the column space


R(A) = Col(A)

(2) Column and row spaces have the same dimension


dimF Col(A) = dimF R(A) := rank(A)
which is called the rank of A.
(3) There is a dimension theorem
rank(A) + nullity(A) = n
where n is the number of columns in A.

14
Theorem 5.2 (L.I. vs L.D.).
In previous theorem, suppose that m = n so that A is a square matrix of order n. Then the
following are equivalent.
(1) A is an invertible matrix, i.e. A has a so called inverse A−1 ∈ Mn (F) such that
AA−1 = In = A−1 A

(2) A has nonzero determinant


det(A) = |A| =
6 0

(3) The column vectors


c1 , . . . , cn
of A form a basis of the column vector n-space Fnc .
(4) The row vectors
r1 , . . . , rn
of A form a basis of the row vector n-space Fnc .
(5) The column vectors
c1 , . . . , cn
of A are linearly indepedent in Fnc .
(6) The row vectors
r1 , . . . , rn
of A are linearly indepedent in Fnc .
(7) The matrix equation
AX = 0
has the trivial solution only: X = 0.
Theorem 5.3 (Row operation preserves columns relations).
Suppose that A and B are row equivalent. Then we have:
(1) If the column vectors
ai1 , . . . , ais , aj1 , . . . , ajt
of A satisfies a relation
ci1 ai1 + · · · + cis ais = cj1 aj1 + · · · + cjt ajt
for some scalars cik ∈ F, then the corresponding column vectors
bi1 , . . . , bis , bj1 , . . . , bjt
of B satisfies exactly the same relation
ci1 bi1 + · · · + cis bis = cj1 bj1 + · · · + cjt bjt
The converse is also true.

15
(2) The column vectors
ai1 , . . . , ais , aj1 , . . . , ajt
of A are linearly dependent if and only if the corresponding column vectors

bi1 , . . . , bis , bj1 , . . . , bjt

of B are linearly dependent.

(3) The column vectors


ai1 , . . . , ais , aj1 , . . . , ajt
of A are linearly independent if and only if the corresponding column vectors

bi1 , . . . , bis , bj1 , . . . , bjt

of B are linearly independent.

(4) The column vectors


ai1 , . . . , ais
of A forms a basis of the column space

Col(A) = Span{a1 , . . . , an }

if and only if the corresponding column vectors

bi1 , . . . , bis

form a basis of the column space

Col(B) = Span{b1 , . . . , bn }

(5) If
B1 := {ai1 , . . . , ais }
is a maximal L.I. subset of the set

C = {a1 , . . . , an }

of all column vectors then B1 is a basis of the column space Col(A) of A.

(6) Suppose that B is in row-echelon form with leading entries at columns

i1 , . . . , i s

Then
ai1 , . . . , ais
forms a basis of the column space Col(A) of A.

16
(7) The row space of A and B are identical

R(A) = R(B)

But the column spaces of A and B may not be the same.

(8) Suppose that 


b01
B = (bij ) =  ... 
 
b0m
is in row-echelon form with leading entries at rows

j1 , . . . t

Then
b0j1 , . . . , b0jt
form a basis of the row space R(A) = R(B) of A.

17
6 Quotient Spaces and Linear Transformations
Definition 6.1 (Sum of subsets of a space).
Let V be a vector space over a field F, and let S and T be subsets (which are not necessarily
subspaces) of V . Define the sum of S and T as

S + T := {s + t | s ∈ S, t ∈ T }

In general, given subsets Si (1 ≤ i ≤ r) of V , we can define the sum of Si as


r
X Xr
Si = { xi | xi ∈ Si }
i=1 i=1

Theorem 6.1 (Inclusion and sum for subsets). (1) Associativity

(S1 + S2 ) + S3 = S1 + (S2 + S3 )

(2) Commutativity
S1 + S2 = S2 + S1

(3) If S1 ⊆ S2 and T1 ⊆ T2 then


S1 + T1 ⊆ S2 + T2

(4) If W is a subspace of V , then

W + {0} = W, W +W =W

(5) Suppose that W is a subspace of V , THen

S+W =W ⇔S ⊆W

Definition 6.2 (Coset v̄).


Let V be a vector space over a field F and W a subspace of V . For any given v ∈ V , the
subset
v + W := {v + w | w ∈ W }
of V is called the coset of W containing v. This subet is often denoted as

v̄ := v + W

The vector v is a representative of the coset v̄.

Theorem 6.2 (Coset Relations).


Let W be a subspace of a vector space V . The following are equivalent

(1) v + W = W ,i.e.,v̄ = 0̄

(2) v ∈ W

18
(3) v + W ⊆ W

(4) W ⊆ v + W

Theorem 6.3 (To be the same coset).


Let W be a subspace of V . Then for v̄i = vi + W ,

v¯1 = v¯2 ⇔ v1 − v2 ∈ W

Remark:Suppose that V = U ⊕ W is a direct sum of subspaces U and W . Then the map


below is a bijection(and indeed an isomorphism)

f : U → V /W
u 7→ ū = u + W

Definition 6.3 (Quotient Space).


Let W be a subspace of V . Let

V /W := {v̄ = v + W | v ∈ V }

be the set of all cosets of W . It is called the quotient space of V modulo W .


We define a binary addition operation on V /W :

+ : V /W × V /W → V /W
¯ v2
(v¯1 , v¯2 ) 7→ v¯1 + v¯2 := v1 +

and a scalar multiplication operation

× : F × V /W → V /W
(a, v¯1 ) 7→ av¯1 := av
¯1

Theorem 6.4 (Quotient Space being Well Defined).


Let V be a vector space over a field F and W a vector subspace of V . Then we have:

(1) The binary addition operation and scalar multiplication operation on V /W is well
defined.

(2) V /W together with these binary addition and scalar multiplication operations, becomes
a vector space over the same field F, with the zero vector

0V /W = 0¯V = w̄

for any w ∈ W .

Definition 6.4 (Linear transformation, and its Kernel and Image; Isomorphism).
Let Vi be two vector spaces over the same field F. A map

ϕ : V1 → V2

19
is called a linear transformation from V1 to V2 if ϕ is compatible with the vector addition
and scalar multiplication on V1 and V2 in the sense below:
ϕ(v1 + v2 ) = ϕ(v1 ) + ϕ(v2 )
ϕ(av) = aϕ(v)
When ϕ : V → V is a linear transformation from V to itself, we call ϕ a linear operator on
V.
A linear transformation is called an isomorphism if it is a bijection. In this case, we denote
V1 ' V2
Remark: If T : V → W is a linear transformation, then T (0V ) = 0W .
Remark:(Direct sum vs quotient space)
Let V = U ⊕ W , where U, W are subspaces of V . Then the map below is an isomorphism.
f : U → V /W
u 7→ ū = u + W
Theorem 6.5 (Equivalent Linear Transformation definition).
Let ϕ : V1 → V2 be a map between two vector spaces Vi over the same field F. The the
following are equivalent.
(1) ϕ is a linear transformation.
(2) ϕ is compatible with taking linear combination in the sense below:
ϕ(a1 v1 + a2 v2 ) = a1 ϕ(v1 ) + a2 ϕ(v2 )
for all ai ∈ F, vi ∈ V .
Theorem 6.6 (Evaluate T at a basis).
Let V be a vector space over a field F and with a basis B = {u1 , u2 , . . .}. Let T : V → W
be a linear transformation.
Then T is uniquely determined by its valuations T (ui ) (i = 1, 2, . . .) at the basis B.
Namely, if T 0 : V → W is another linear transformation such that T 0 (ui ) = T (ui )∀i, then
they are equal: T 0 = T .
Theorem 6.7 (Quotient map).
Let V be a vector space over a field F. Let W be a subspace V and V /W . One verifies that
γ is surjective and
ker(γ) = W
Theorem 6.8 (Image being a vector subspace).
Let
ϕ:V →W
be a linear transformation between two vector spaces over the same field F. Let V1 be a
vector subspace of V . Then the image of V1 :
T (V1 ) = {T (u | u ∈ V1 }
is a vector subspace of W .
In particular, T (V ) is a vector subspace of W .

20
Theorem 6.9 (Subspace vs. Kernel).
Let V be a vector space over a field F.
(1) Suppose that
ϕ:V →U
is a linear transformation. Then the kernel ker(ϕ) is a vector subspace of V .
(2) Conversely, suppose W is a vector subspace of V . Then there is a linear transformation

ϕ:V →U

such that
W = ker(ϕ)

Theorem 6.10 (To be injective). Let

ϕ:V →W

be a linear transformation. Show that ϕ is injective if and only if ker(ϕ) = {0}.


Theorem 6.11 (Equivalent Isomorphism Definition).
Let ϕ : V → W be a linear transformation. Then there is an isomorphism
ϕ̄ : V / ker(ϕ) ' ϕ(V ) ∈ U
v̄ 7→ ϕ(v)
such that
ϕ = ϕ̄ ◦ γ
where
γ : V → V / ker(ϕ)v 7→ v̄
is the quotient map, a linear transformation.
In particular, when ϕ is surjective, we have an isomorphism

ϕ̄ : V / ker(ϕ) ' U

Theorem 6.12 (Finding basis of the quotient).


Let V be a vector space over a field F of finite dimension n. Let W be a subspace with a
basis B1 = {w1 , . . . , wr }.
(1) B1 extends to a basis a
B := B1 {w1 , . . . , wr }
of V .
(2) The cosets
{w̄1 , . . . , w̄r }
is a basis of the quotient space V /W . In particular,

dimF V /W = dimF V − dimF W

21
(3) a
B1 {ur+1 , . . . , un }
is a basis of V if and only if the cosets

{ur+1
¯ , . . . , u¯n }

is a basis of V /W .
Theorem 6.13 (Goodies of Isomorphism).
Let ϕ : V → W be an isomorphism and let B be a subset of V . Then we have:
(1) If there is a relation
r
X s
X
ai vi = a i vi
i=1 i=r+1

among vectors vi ∈ V , then exactly the same relation


r
X s
X
ai ϕ(vi ) = ai ϕ(vi )
i=1 i=r+1

holds among vectors ϕ(vi ) ∈ W . The converse is also true.

(2) B is linearly dependent if and only if so is ϕ(B).

(3) B is linearly independent if and only if so is ϕ(B);

(4) We have
ϕ(Span(B)) = Span(ϕ(B))
In particular,
ϕ(Span{v1 , . . . , vs }) = Span{ϕ(v1 ), . . . , ϕ(vs )}

(5) B spans V if and only if ϕ(B) spans W .

(6) B is a basis of V if and only if ϕ(B) is a basis of W . In particular,

dim V = dim W

Theorem 6.14 (To be isomorphic finite-dimensional spaces).


Let V and W be finite-dimensional vector spaces over the same field F. Then the following
are equivalent.
(1) dimF V = dimF W = n.

(2) There is an isomorphism


ϕ:V 'W

(3) For some n, we have:


V ' Fn ' W

22
Theorem 6.15 (Dimension Theorem).
Let ϕ : V → W be a linear transformation between vector spaces over a field F. Then

dimF ker(ϕ) + dimF ϕ(V ) = dimF V

Theorem 6.16 (2nd Isomorphism Theorem).


Let W1 , W2 be vector subspaces of a vector space V .

(1) The map


ϕ : W + 1/(W1 + W2 )) → (W1 + W2 )/W2
w + W1 ∩ W2 = w 7→ (w) = w + W2
is a well difined isomorphism between vector spaces.

(2) A dimension formula:

dim W1 + dim W2 = dim(W1 + W2 ) + dim(W1 ∩ W2 )

Theorem 6.17 (Equivalent isomorphism definition).


Let
ϕ:V →W
be a linear transformation between vector spaces over a field F and of the same finite dimen-
sion n. Then the following are equivalent.

(1) ϕ is an isomorphism

(2) ϕ is an injection

(3) ϕ is an surjection

23
7 Representation Matrices of Linear Transformations
Definition 7.1 (Coordinate vector).
Let V be vector space of dimension n ≥ 1 over a field F. Let
B = BV = (v1 , . . . , vn )
be a basis of V . Every vector v ∈ V can be expressed as a linear combination
v = c1 v1 + · · · + cn vn
and this expression is unique. We gather the coefficients ci and form a column vector
 
c1
 .. 
[v]B :=  .  ∈ Fnc
cn
which is called the coordinate vector of v related to basis B.
One can recover v from its coordinate vector [v]B :
v = B[v]B
Theorem 7.1 (Isomorphism V → Fnc ).
Let V be an n-dimensional vector space over a field F and with a basis B = {v1 , . . . , vn }.
Show that the map
ϕ : V → Fnc
v 7→ [v]B
is an isomorphism between the vector space V and Fnc .
Theorem 7.2 (Representation matrix).
Let
T :V →W
be a linear transformation between vector spaces over a field F. Let
B = {v1 , . . . , vn }
be a basis of V , and
BW = {w1 , . . . , wm }
be a basis of W . Let A ∈ Mm×n (F). Then the following three conditions on A are equivalent.
[T (v)]BW = A[v]B
A = ([T (v1 )]BW , . . . , [T (vn )]BW )
(T (v1 ), . . . , T (vn )) = (w1 , . . . , wm )A
We denote the above matrix A as
[T ]B,BW := A = ([T (v1 )]BW , . . . , [T (vn )]BW )
and call it the representation matrix of T relative to B and BW .

24
Theorem 7.3 (Linear transformation theory = Matrix theory).
For every matrix
A ∈ Mm×n (F)
there is a unique linear transformation
T :V →W
such that the representation matrix
[T ]B,BW = A
Consequently, the map
ϕ : HomF (V, W ) → Mm×n (F)
T 7→ [T ]B,BW
is an isomorphism of vector spaces over F.
Theorem 7.4 (Close relation between the space of vectors and space of their coordinates).
Let
T :V →W
be a linear transformation between the vector spaces V and W over the same field F, of
dimensions n and m, respectively. Let
BV := (v1 , . . . , vn )
be a basis of V , and
BW := (w1 , . . . , wm )
be a basis of W . Let
Mm×n (F) 3 A := [T ]BV ,BW = (a1 , . . . , an )
Then the following are isomorphisms:
ϕ : Ker(T ) → Null(A)
v 7→ [v]BV ,
ψ : Null(A) → Ker(T )
X 7→ BV X,
ξ : R(T ) → R(TA ) = col.sp. of A = Span{a1 , . . . , an }
η : R(TA ) → R(T )
Y 7→ BW Y = (w1 , . . . , wm )Y
Below are some consequences of the isomorphism above
(1) The subset
{X1 , . . . , Xs }
of Fnc is a basis of Null(A) if and only if the vectors
BV X1 , . . . , BV XS
of V forms a basis of Ker(T ).

25
(2) The subset
{Y1 , . . . , Yt }
of Fm
c is a basis of R(TA ) if and only if the vectors

BW Y1 , . . . , BW Yt

of W form a basis of R(T ).

(3) The range of T is given by

R(T ) = Span{BW a1 , . . . , BW an }

(4) T : V → W is an isomorphism if and only if its representation matrix A = [T ]BV ,BW is


an invertible matrix in Mn (F).

Theorem 7.5 (Representation Matrix of a Composite Map).


Let V1 , V2 , V3 be vector spaces of finite dimension over the same field F and let B1 , B2 , B3 be
theire respective bases. Let
T1 : V1 → V2
and
T2 : V2 → V3
be linear transformations. Then we have:

[T2 ◦ T1 ]B1 ,B3 = [T2 ]B2 ,B3 [T1 ]B1 ,B2

Theorem 7.6 (Representation matrix of inverse of an isomorphism).


Let
T :V →W
be an isomorphism between vector spaces over the same field F and of finite dimension. Let

T −1 : W → V

be the inverse isomorphism of T . Let BV (resp. BW ) be a basis of V (resp. W ). Then

[T −1 ]BW ,BV = [TBV ,BW ]−1

Theorem 7.7 (Representation matrix of map combination).


Let
Ti : V → W
be two linear transformations between finite-dimensional vector spaces over the same field F.
Let B(resp. BW ) be a basis of V (resp. W ). Then for any ai ∈ F, the map linear combination
a1 T1 + a2 T2 has the representation matrix

[a1 T1 + a2 T2 ]B,BW = a1 [T1 ]B,BW + a2 [T2 ]B,BW

26
Theorem 7.8 (Equivalent transition matrix definition).
Let V be a vector space over a field F and of finite dimension n ≥ 1. Let

B = (v1 , . . . , vn )

and
B 0 := (v10 , . . . , vn0 )
be two bases of V . Let P ∈ Mn (F). Then the following are equivalent.

(1)
P = ([v10 ]B , . . . , [vn0 ]B )

(2)
B 0 = BP

(3) For any v ∈ V , we have


P [v]B 0 = [v]B

This P is denoted as PB 0 →B and called the transition matrix from basis B 0 to B. P is


invertible.

Theorem 7.9 (Basis change theorem for representation matrix).


Let V be a vector space over a field F and of finite dimension n ≥ 1. Let

B = (v1 , . . . , vn )

and
B 0 := (v10 , . . . , vn0 )
be two bases of V . Then
[T ]B 0 = P −1 [T ]B P
where
P = PB 0 →B

Definition 7.2 (Similar Matrices).


Two square matrices (of the same order) A1 , A2 ∈ Mn (F) are similar if there is an invertible
matrix P ∈ Mn (F) such that
A2 = P −1 A1 P
In this case, we denote
A1 ∼ A2
The similarity property is an equivalence relation.

Theorem 7.10. Similar matrices have the same determinant:

A1 ∼ A2 ⇒ |A1 | = |A2 |

27
Definition 7.3 (Determinant/Trace of a linear operator).
Let
T :V →V
be a linear operator on a finite-dimensional vector space V .
We define the determinant det(T ) of T as

det(T ) := det([T ]B )

and the trace of T as


Tr(T ) = Tr([T ]B )
where B is any basis of V .

Definition 7.4 (Characteristic polynomial pA (x), pT (x)). (1) Let A ∈ Mn (F).

pA (x) : = |xIn − A|
= xn + bn−1 xn−1 + · · · + b1 x + b0

is called the characteristic polynomial of A, which is of degree n.

(2) Let
T :V →V
be a linear operator on an n-dimensional vector space V . Set

A := [T ]B

where B is any basis of V . Then

pT (x) : = |xIn − A|
= xn + bn−1 xn−1 + · · · + b1 x + b0

is called the characteristic polynomial of T , which is of degree n = dim V .

Theorem 7.11. Similar matrices have equal characteristic polynomial.

Theorem 7.12.
For A ∈ Mn (F), we have
Tr(A) = −bn−1
det(A) = (−1)n pA (0)

28
8 Eigenvalue and Cayley-Hamilton Theorem
Definition 8.1 (Eigenvalue, eigenvector).
Assume that
λ∈F

(arabic*) Let V be a vector space over a field F. Let

T :V →V

be a linear operator. A nonzero vector v in V is called an eigenvector of T corre-


sponding to the eigenvalue λ ∈ F of T if

T (v) = λv

(arabic*) For an n × n matrix A in Mn (F), a nonzero column vector u in Fnc is called an eigen-
vector of A corresponding to the eigenvalue λ ∈ F of A if

Au = λu

Definition 8.2 (Equivalent definition of eigenvalue and eigenvector).


Let V be a vector space of dimension n over a field F and with a basis B, Let

T :V →V

be a linear operator. Assume that


λ∈F
Then the following are equivalent:

(1) λ is an eigenvalue of T (corresponding to an eigenvector 0 6= v ∈ V of T , i.e. T (v) =


λv).

(2) λ is an eigenvalue of [T ]B (corresponding to an eigenvector 0 6= [v]B ∈ Fnc of [T ]B , i.e.


[T ]B [v]B = λ[v]B ).

(3) The linear operator


λIV − T : V → V
x 7→ λx − T (x)
is not an isomorphism, i.e. there is some

0 6= v ∈ Ker(λIV − T )

(4) The matrix λIn − [T ]B is not invertiblem i.e. the matrix equation

(λIn − [T ]B )X = 0

has a non=trivial solution.

29
(5) λ is a zero of the characteristic polynomial pT (x) of T

pT (λ) = |λIn − [T ]B | = 0

Theorem 8.1 (Determinant |A| as product of eigenvalues).


Let A ∈ Mn (F). Let p(x) be the characteristic polynomial. Factorise

p(x) = (x − λ1 ) · · · (x − λn )

in some over field of F. Then the determinant of A equals


n
Y
λi
i=1

Definition 8.3 (Eigenspace of an eigenvalue).


Let λ ∈ F be an eigenvalue of a linear operator

T :V →V

on an n-dimensional vector space V over the field F. The subspace (of all the eigenvectors
corresponding to the eigenvalue λ, plus 0V ):

Vλ : = Vλ (T )
: = Ker(λIV − T )
= {v ∈ V | T (v) = λv}

of V is called the eigenspace of T corresponding to the eigenvalue λ.

Definition 8.4 (Geometric/Algebraic Multiplicity).


Let λ ∈ F and T : V → V be as the previous definition.

(1) The dimension


dim Vλ
of the eigenspace Vλ of T is called the geometrix multiplicity of the eigenvalue λ of
T . We have
1 ≤ dim Vλ ≤ n

(2) The algebraic multiplicity of the eigenvalue λ of T is defined to be the largest positive
integer k such that (x − λ)k is a factor of the characteristic polynomial pT (x),i.e.

(x − λ)k | pT (x), (x − λ)k+1 - pT (x)

We shall see that

geometric multiplicity of λ ≤ alg. multiplicity of λ

30
Theorem 8.2 (Eigenspace of T and [T ]B ).
Let
T :V →V
be a linear operator on an n-dimensional vector space V with a basis B. Set

A := [T ]B

The map
f : Ker(T − λIV ) → Null(A − λIn )
w 7→ [w]B
gives an isomorphism. In particular,

dim Vl ambda(T ) = dim Vλ (A)

Due to this isomorphism, the following are equivalent:

1. The subset
{u1 , · · · , us }
of V is a basis of the eigenspace Vλ (T ) of T .

2. THe subset
{[u1 ]B , cdots, [us ]B }
of Fnc is a basis of the eigenspace Vλ ([T ]B ) of the representation matrix [T ]B of T
relative to a basis B of V .

Also, the following are equivalent:

1. The subset
{X1 , · · · , Xs }
of Fnc is a basis of the eigenspace Vλ ([T ]B ) of the representation matrix [T ]B of T
relative to a basis B of V .

2. The subset
{BX1 , · · · , BXs }
of V is a basis of the eigenspace Vλ (T ) of T .

Theorem 8.3 (Eigenspaces of similar matrices).


Let A ∈ Mn (F). Suppose that P −1 AP = C. Show that

Fnc ⊇ Vλ (A) = P Vλ (C) := {P X | X ∈ Vλ (C)}

Theorem 8.4 (Sum of eigenspaces).


Let
λ1 , . . . , λ k

31
be some distinct eigenvalues of a linear operator T on a vector space V over a field F. Then
the sum of eigenspaces
Xk
W := Vλi (T )
i=1
= Vλi (T ) + · · · + Vλk (T )
is a direct sum:
W = ⊕ki=1 Vλi (T )
= Vλ1 (T ) ⊕ · · · ⊕ Vλk (T )

Definition 8.5 (Multiplication of linear operators S1 , . . . , Sr ).


Let
T :V →V
be a linear operator on a vector space V over a field F. Define

T s := T ◦ · · · ◦ T (s times)

which is a linear operator on V .


Ts : V → V
v 7→ T s (v)
Be convention, set
T 0 := IV = idV
More generally, for a polynomial
r
X
f (x) = ai xi
i=0

Define r
X
f (T ) = ai T i
i=0

Then f (T ) is a linear operator on V .

f (T ) : V &toV
v 7→ f (T )(v)

Similarly, for linear operators


Si : V → V
Define
S1 S2 · · · Sr := S1 ◦ S2 ◦ · · · ◦ Sr
which is a linear operator.

Theorem 8.5 (Polynomials in T ).


Let
T :V →V

32
be a linear operator on a vector space V over a field F and with a basis

B = {u1 , . . . , un }

Let
f (x), g(x) ∈ F[x]
be polynomials. We have

(1)
[f (T )]B = f ([T ]B )

(2) The multiplication f (T )g(T ) as polynomials in T equals the composite f (T ) ◦ g(T ) as


linear operators:
f (T )g(T ) = f (T ) ◦ g(T )

(3) Commutativity:
f (T )g(T ) = g(T )f (T )

(4) If P ∈ Mn (F) is invertible, then

f (P −1 AP ) = P −1 f (A)P

(5) If
S:V →V
is an isomorphism with inverse isomorphism

S −1 : V → V

Then
f (S −1 T S) = S −1 f (T )S

Definition 8.6 (T -invariant Subspace).


Let
T :V →V
be a linear operator on a vector space V . A subspace W of V is called T -invariant if the
image of W under the map T is included in W :

T (W ) := {T (w) | w ∈ W }

i.e.,
T (w) ∈ W ∀w ∈ W
In this case, define the restriction of T on W as:

T |W :W →W
w 7→ T (w)

33
Theorem 8.6 (Kernels and Images of Commutative Operators).
Let Ti : V → V be two linear operator commutative to each other, i.e.
T1 ◦ T2 = T2 ◦ T1
as maps. Namely,
T1 (T2 (v)) = T2 (T1 (v)) (∀v ∈ V )
Both
Ker(T2 ), ima(T2 )
are T1 -invariant subspaces of V .
Theorem 8.7 (Evaluate T on a basis of a subspace).
Let
T :V →V
be a linear operator on a vector space V over a field F. Let W be a subspace of V with a
basis BW = {w1 , w2 , · · · }. Then W is T -invariant, if and only if
T (BW ) ⊆ W
Theorem 8.8 (T -cyclic subspace).
Let
T :V →V
be a linear operator on a vector space V over a field F. Fix a vector
0 6= w1 ∈ V
1. The subspace
W := Span{T s (w1 | s ≥ 0}
of V is T -invariant.
W is called the T -cyclic subspace of V generated by w1 .
2. Suppose that V is finite-dimensional. Let s be the smallest positive integer such that
T s (w1 ) ∈ Span{w1 , T (w1 ), . . . , T s−1 (w1 )}
We have
dimF W = s
and
B := {w1 , T (w1 ), . . . , T s−1 (w1 )}
is a basis of W .
3. In (2), if
T s (w1 ) = c0 w1 + c1 T (w1 ) + · · · + cs−1 T s−1 (w1 )
for some scalars ci ∈ F, then the characteristic polynomial of the restiction operator
T | W on W is
pT |W (x) = −c0 − c1 x − · · · − cs−1 xs−1 + xs

34
Theorem 8.9 (Characteristic Polynomial of the Restriction Operator).
Let
T :V →V
be a linear operator on a vector space V over a field F and of dimension n ≥ 1. Let W be
a T -invariant subspace of V . Then the characteristic polynomial pT |W (x) of the restriction
operator T | W on W is a factor of the characteristic polynomial pT (x) of T , i.e.

pT (x) = q(x)pT |W (x)

for some polynomial q(x) ∈ F[x].


Theorem 8.10 (To be T -invariant in terms of [T ]B ).
A subspace W of an n-dimensional space V is T -invariant for a linear operator T on V , if
and only if every basis BW of W can be extended to a basis

B = BW ∪ B2

of V such that the representation matrix of T relative to B, is of the form:


 
A1 A2
[T ]B =
0 A3

for some square matrices A1 , A3 (automatically with A1 = [T | W ]BW ).


In this case, the matrix
A2 = 0
if and only if
W2 := Span(B2 )
is a T -invariant subspace of V (automatically with [T | W2 ]B2 = A3 )
Theorem 8.11 (Upper Triangular Form of a Matrix).
Let T : V → V be a linear operator on an n-dimensional vector space V over a field F.
Suppose that the characteristic polynomial p(x) is factorised as

p(x) = (x − λ1 )n1 · · · (x − λk )nk

for some λi ∈ F. Then there is an basis B of V such that the representation matrix [T ]B is
upper triangular.
Theorem 8.12 (Characteristic Polynomials of Direct Sums).
Let
T :V →V
be a linear operator on an n-dimensional vector space V over a field F. Suppose that there
are T -invariant subspaces
Wi (1 ≤ i ≤ r)
of V such that V is the direct sum
V = ⊕ri=1 Wi

35
of Wi .
Then the characteristic polynomial pT (x) of T is the product:
r
Y
pT (x) = pT |Wi (x)
i=1

of the characteristic polynomials of the restriction operators T | Wi on Wi .

Theorem 8.13 (To be direct sum of T -invariant subspaces).


An n-dimensional vector space V with a linear operator T is a direct sum

V = ⊕ri=1 Wi

of some T -invariant subspaces Wi , if and only if every set of bases Bi of Wi gives rises to a
basis a a
B = B1 ··· Br
of V such that the representation matrix of T relative to B is in the form
 
A1 0 · · · 0
 0 A2 · · · 0
[T ]B =  ..
 
.. . . .. 
 . . . . 
0 0 0 Ar

with Ai of order |Bi | = dim Wi

Theorem 8.14 (Cayley-Hamilton Theorem).


Let n
X
pT (x) = |xIn − [T ]B | = bi x i
i=0

be the characteristic polynomial of a linear operator

T :V →V

on an n-dimensional vector space V over a field F and with a basis B. Then T satisfies the
equation pT (x) = 0, i.e.
pT (T ) = 0IV
which is the zero map on V .

36
9 Minimal Polynomial and Jordan Canonical Form
Definition 9.1 (Minimal Polynomial).
Let
T :V →V
be a linear operator on an n-dimensional vector space over a field F. A nonzero polynomial
m(x) ∈ F[x]
is a minimal polynomial of T if it satisfies:
1. m(x) is monic,
2. Vanishing condition:
m(T ) = 0IV
3. Minimality degree condition:
Whenever f (x) ∈ F[x] is another nonzero polynomial such that f (T ) = 0IV , we have
deg(f (x)) ≥ deg(m(x))

We can define a minimal polynomial of a matrix A ∈ MN (F).


Remark: The existence of minimal polynomial is proven by Cayley-Hamilton theorem.
Theorem 9.1 (Uniqueness of a minimal polynomial mT (x)).
Let T : V → V be a linear operator on an n-dimensional vector space V over a field F. Let
m(x) be a minimal polynomial of T . Let f (x) ∈ F[x]. Then the following are equivalent.
(1) f (T ) = 0IV .
(2) m(x) is a factor of f (x),i.e., m(x) | f (x).
In particular, there is exactly one minimal polynomial of T and will be denoted as
mT (x) = m(x)
Further, if A = [T ]B , then mT (x) = mA (x).
Theorem 9.2 (Minimal polynomials of similar matrices).
If two matrices Ai are simialr: A1 ∼ A2 , then they have the same minimal polynomial
mA1 (x) = mA2 (x)
Theorem 9.3 (Minimal polynomials of direct sums).
Consider the matrix  
A1 0 · · · 0
 0 A2 · · · 0 
A =  ..
 
.. . . .. 
 . . . . 
0 0 · · · Ar
where Ai ∈ Mni (F) are square matrices. The minimal polynomial mA (x) of A is equal to
the least common multiple of the minimal polynomials mAi (x) of Ai ,i.e.,
mA (x) = lcm{mA1 (x), . . . , mAr (x)}

37
Theorem 9.4.
The set of zeros of pT (x) and that of mT (x) are identical.
Definition 9.2 (Jordan Block).
Let λ be a scalar in a field F. The matrix below
 
λ 1 0 0 ···
0
 λ 1 0 ···

J := Js (λ) =  0
 0 λ 0 ···
 ∈ Ms (F)
 .. .. .... 
..
. . . . .
0 0 0 ··· λ

is called the Jordan Block of order s with eigenvalue λ.


The characteristic polynomial and minimal polynomial of J are identical:

mJ (x) = (x − λ)s = pJ (x)

The eigenspace
Vλ (J) = Span{e1 }
has dimension 1, i.e. the geometric multiplicity of λ is 1, but the algebraic multiplicity of λ
is s.
Definition 9.3 (Jordan Canonical Form).
Let λ be a nonzero scalar in a field F. Let

s1 ≤ s2 ≤ · · · ≤ se

The following Block Diagonal


 
Js1 (λ) 0 0 ··· 0
 0
 Js2 (λ) 0 ··· 0 

 0
A(λ) =  0 Js3 (λ) ··· 0 

 .. .. .. .. .. 
 . . . . . 
0 0 0 · · · Jse (λ)

is called a Jordan canonical form with eigenvalue λ.


The order of A(λ) is
Xe
s= si
i=1

The characteristic polynomial and minimal polynomial of A are

pA(λ) (x) = (x − λ)s , mA(λ) (x) = (x − λ)se

where s is also called algebraic multiplicity of λ of A(λ).


The eigenspace of A

Vλ (A(λ)) = Span{e1 , e1+s1 , . . . , e1+s1 +···+se−1 }

38
has dimension equal to e.
The geometric multiplicity of the eigenvalue λ of A(λ) is dim Vλ (A(λ)) = e. And we have,

e≤s

More generally, let


λ1 , . . . , λ k
be distinct scalars in F. WLOG, we assume λ1 < λ2 < · · · < λk when F = R. Then the
block diagonal  
A(λ1 ) 0 0 ··· 0
 0
 A(λ2 ) 0 ··· 0  
J = 0
 0 A(λ3 ) · · · 0  
 .. .. .. . . .. 
 . . . . . 
0 0 0 · · · A(λk )
is called a Jordan canonical form where A(λi ) is a Jordan canonical form with eigenvalue
λi as shown above.
Each A(λi ) is of order
s(λi )
and s(λi ) is also the number of times the same scalar λi appears on the diagonal of J and
also the algebraic multiplicity of the eigenvalue λi of J.
So J is of order euqal to
Xk
s(λi )
i=1

There are exactly


e(λi )
Jordan blocks (with eigenvalue λi ) in A(λi ), the largest of which is of order

se (λi )

This se (λi ) is also the multiplicity of λi in the minimal polynomial mJ (x).


There are exactly
Xk
e(λi )
i=1

Jordan blocks in J.
Now we have
k
Y
pJ (x) = (x − λi )s(λi )
i=1
Yk
mJ (x) = (x − λi )se (λi )
i=1

39
The eigenspace Vλi (J) has dimension e(λi ) and is spanned by the e(λi ) ectors corresponding
to the first columns of the e(λi ) Jordan blocks in A(λi ).
We also have
dim Vλi (J) = e(λi ) ≤ s(λi )
Sometimes, a block diagonal J below
 
Js1 (λ1 ) 0 0 ··· 0
 0
 Js2 (λ2 ) 0 ··· 0 

J =
 0 0 Js3 (λ3 ) ··· 0 

 .. .. .. ... .. 
 . . . . 
0 0 0 · · · Jsr (λr )

is also called a Jordan Canonical Form, where each Jsi (λi ) is a Jordan block with eigen-
value λi ∈ F, but these λi ’s may not be distinct.
Assume that there are exactly k distinct elements in the set

{λ1 , . . . , λr }

and we assume that


λmi
are these k distinct ones. These k of λmi are just the distinct eigenvalues of J.
Let
s(λmi )
be the number of times the same scalar λmi appears on the diagonal of J. Let

e(λmi )

be the number of Jordan blocks (among the r such in J) with eigenvalue of the same λmi ;
among these e(λmi ) Jordan blocks, the largest is of order say

se (λmi )

The eigenspace Vλmi (J) has dimension e(λmi ) and is spanned by e(λmi ) vectors corresponding
to the first columns of these e(λmi ) Jordan blocks.
Also,
Yk
pJ (x) = (x − λi )s(λmi )
i=1
Yk
mJ (x) = (x − λi )se (λmi )
i=1

As in the case of A(λ), for the matrix J, we have

dim Vλmi (J) = e(λmi ) ≤ s(λmi )

40
Theorem 9.5 (Jordan Canonical Form of a Linear Operator).
Let V be a vector space of dimension n over a field F and

T :V →V

a linear operator with characteristic polynomial pT (x) and minimal polynomial mT (x) as
follows
pT (x) = (x − λ1 )n1 · · · (x − λk )nk
mT (x) = (x − λ1 )m1 · · · (x − λk )mk
where
λ1 , . . . , λ k
are distinct scalars in F.
Then there is a basis B of V such that the representative matrix [T ]B equals a Jordan
canonical form J ∈ Mn (F), with

ni = s(λi ), mi = se (λi )

Such a block diagonal J is called a Jordan canonical form of T . It is unique up to re-ordering


of λi . The basis B of V is called a Jordan canonical basis of T .

Theorem 9.6 (A canonical form of T is a canonical form of [T ]B ).


Let V be an n-dimensional vector space over a field F and

T :V →V

a linear operator. Let


A = [T ]0B
be the representation matrix of T relative to a basis

B 0 = (v1 , . . . , vn )

of V . Let J be a Jordan canonical form. Then the following are equivalent:

(1) There is an invertible matrix P ∈ Mn (F) such that

P −1 AP = J

(2) THere is an invertible matrix P ∈ Mn (F) such that the representation matrix [T ]B
relative to the new basis
B = B0P
is J, i.e.
[T ]B = J

Theorem 9.7 (Existence of Jordan Canonical Form).


Let F be a field. Let A be a matrix in Mn (F). Let p(x) = pA (x) be the characteristic
polynomial. Then the following is equivalent.

41
(1) A has a Jordan canonical form J ∈ Mn (F).

(2) Every zero of the characteristic polynomial p(x) belongs to F.

(3) We can factor p(x) as


p(x) = (x − λ1 ) · · · (x − λn )
where all λi ∈ F.

In particular, if F is so called algebraically closed, then every matrix A ∈ Mn (F) and every
T on an n-dimensional vector space V over F have a Jordan canonical form j ∈ Mn (F).

Theorem 9.8 (Consequences of Jordan canonical forms).


Let A be a matrix in Mn (F). Set p(x) = pA (x) and m(x) = mA (x).

1. The characteristic polynomial p(x) and the minimal polynomial m(x) have the same
zero sets.
{α ∈ F | p(α) = 0} = {α ∈ F | m(α) = 0}
Also, the multiplicity ni and mi of a zero λi of p(x) and m(x) satisfy

ni ≥ mi ≥ 1

2. If J ∈ Mn (F) is a Jordan canonical form of A, then we have

dim Vλi (A) = dim Vλi (J) = e(λi ) ≤ s(λi )

Theorem 9.9 (Canonical forms of similar matrices).


Let Ai ∈ Mn (F) and Ji ∈ Mn (F) be its Jordan canonical form. Then the following are
equivalent.

(1) A1 and A2 are similar.

(2) We have J1 = J2 after re-ordering of their Jordan Block.

Definition 9.4 (Diagonalisable Operator).


Let V be an n-dimensional vector space over a field F. A linear operator T : V → V is
diagonalisable over F, if the representation matrix [T ]B relative to some basis B of V is a
diagonal matrix in Mn (F):
 
λ1 0 0 · · · 0
 0 λ2 0 · · · 0 
 
[T ]b = J =  0 0 λ3 · · · 0 
 
 .. .. .. . . .
. . . . .. 
0 0 0 · · · λn

where λi are scalars in F. This J is then an automatically a Jordan canonical form of T .


Clearly,
λ1 , . . . , λn

42
exhaust all zeros of pT (x) and the characteristic polynomial of T is

pT (x) = (x − λ1 ) · · · (x − λn )

A square matrix A ∈ Mn (F) is diagonalisable over F, if A is similar to a diagonal matrix in


Mn (F), i.e.  
λ1 0 0 ··· 0
 0 λ2
 0 ··· 0 
−1
P AP = J =  0 0
 λ3 ··· 0 
 .. .. .. .. .. 
. . . . .
0 0 0 · · · λn
for some invertible P ∈ Mn (F), where λi are scalars in F. This J is then automatically a
Jordan canonical form of A.
Write
P = (p1 , . . . , pn )
with pj the jth column of P .
The diagonalisability condition on A is equivalent to
 
λ1 0 0 · · · 0
 0 λ2 0 · · · 0
 
AP = P  0 0 λ3 · · · 0


 .. .. .. . . .. 
. . . . .
0 0 0 · · · λn

i.e.
(Ap1 , . . . , Apn ) = (λ1 p1 , . . . , λn pn )
i.e. each pi is an eigenvector of A corresponding to the eigenvalue λi .
Suppose that A = [T ]B 0 . Then the condition above is equivalent to

[T (vi )]B 0 = [T ]B 0 [vi ]B 0 = λi [vi ]B 0

where vi = B 0 pi ∈ V with
[vi ]B 0 = pi
i.e.
T (vi ) = λi vi
i.e.  
λ1 0 0 ··· 0
 0 λ2 0
 ··· 0 
T (v1 , . . . , vn ) = (v1 , . . . , vn )  0 0 λ3
 ··· 0 
 .. .. .. .. .. 
. . . . .
0 0 0 · · · λn

43
i.e.  
λ1 0 0 ··· 0
 0 λ2 0
 ··· 0 
[T ]B =  0 0 λ3
 ··· 0 
 .. .. .. .. .. 
. . . . .
0 0 0 · · · λn
where
B = (v1 , . . . , vn ) = B 0 P
is a basis of V since
([v1 ]B 0 , . . . , [vn ]B 0 ) = (p1 , . . . , pn )
is a basis of column vector space Fnc .
Theorem 9.10.

(1) T is diagonalisable if and only if the representation matrix [T ]B 0 relative to every basis
B 0 is diagonalisable.

(2) A matrix A ∈ Mn (F) is diagonalisable if and only if the matrix transformation TA on


the column n-space Fnc is diagonalisable.
Theorem 9.11 (Equivalent Diagonalisable Condition).
Let V be an n-dimensional vector space over a field F, and

T :V →V

a linear operator. Then the following are equivalent:


1. T i s diagonalisable over F, i.e. the representation matrix of T relative to some basis
B of V is a diagonal matrix in Mn (F).
 
λ1 0 0 · · · 0
 0 λ2 0 · · · 0 
 
 0 0 λ3 · · · 0 
[T ]B =  
 .. .. .. . . .. 
. . . . .
0 0 0 · · · λn

2. [T ]B 0 is diagonalisable over F for every basis B 0 of V , i.e. there exists an invertible


P ∈ Mn (F) such that
P −1 [T ]B 0 P = diag[λ1 , . . . , λn ]
for some scalars λi ∈ F(automatically being eigenvalues of T ).

3. A basis
B = (v1 , . . . , vn )
of V is formed by eigenvectors vi of T .

44
4. There are n linearly independent eigenvectors vi of T .

5. For the representation matrix [T ]B 0 relative to every basis B 0 of V , a basis

P = (p1 , . . . , pn )

of the column n-space Fnc is formed by eigenvectors pi of [T ]B 0 .

6. For the representation matrix [T ]B 0 relative to every basis B 0 of V , there are n linearly
independent eigenvectors pi of [T ]B 0 .

7. Let
λm1 , . . . , λmk
be the only distinct eigenvalues of T and let Bi be a basis of the eigenspace Vλi (T ).
Then
B = (B1 , . . . , Bk )
is a basis of V , automatically with
 
λm1 I|B1 | 0 0 ··· 0
 0
 λm2 I|B2 | 0 ··· 0 

 0
[T ]B =  0 λm3 I|B3 | ··· 0 

 .. .. .. ... .. 
 . . . . 
0 0 0 · · · λmk I|Bk |

8. Let
λm1 , . . . , λmk
be the only distinct eigenvalues of T . Then V is a direct sum of the eigenspaces

V = Vλmi (T ) ⊕ · · · ⊕ Vλmk (T )

9. Let
λm1 , . . . , λmk
be the only distinct eigenvalues of T . Then
k
X
dim Vλmi (T ) = dim V
i=1

10. T has a Jordan canonical form J which is diagonal.

Theorem 9.12 (Minimal polynomial and diagonalisability).


Let F be a field. Let A be a matrix in Mn (F). Let m(x) = mA (x) be minimal polynomial of
A. Then the following are equivalent:

(1) A is diagonalisable over F.

45
(2) The minimal polynomial m(x) is a product of distinct linear polynomials in F[x].

m(x) = (x − λ1 ) · · · (x − λk )

where λi are distinct scalars in F


(3) We can factor m(x) over F as

m(x) = (x − λ1 ) · · · (x − λk )

for some scalars λi ∈ F and m(x) has only simple zeros.


(4) Let p(x) = pA (x) be the characteristic polynomial. Then we can factorise p(x) over F
as
p(x) = (x − λ1 )n1 · · · (x − λk )nk
where λi are distinct scalars in F. The dimension of the eigenspace satisfies:

dim Vλi = ni

Theorem 9.13.
Let V be an n-dimensional vector space over a field F.
A linear operator
T :V →V
on V is nilpotent if
T m = 0IV
for some positive integer m.
Suppose that T has a Jordan canonical form J ∈ Mn (F). The following are equivalent.
(1) T is nilpotent.
(2) J equals some A(λ) with λ = 0.
(3) Every eigenvalue of T is zero/
(4) The characteristic polynomial of T is pT (x) = xn .
(5) The minimal polynomial of T is mT (x) = xs for some s ≥ 1.
Theorem 9.14 (Additive Jordan Decomposition).
Suppose a linera operator
T :V →V
has a Jordan canonical form in Mn (F). There are linear operators

TS : V → V

and
Tn : V → V
satisfying the following:

46
(1) A decomposition
T = Ts + Tn

(2) Ts is semi-simple.

(3) Tn is nilpotent.

(4) Commutativity
Ts ◦ Tn = Tn ◦ Ts

(5) There are polynomials f (x), g(x) in F[x] such that

Ts = f (T ) Tn = g(T )

This decomposition is unique. We call it Jordan decomposition.

47
10 Quadratic Forms, Inner Product Spaces and Conics
Definition 10.1 (Bilinear forms).
Let V be a vector space over a field F. Consider the map H below:
H :V ×V →F
(x, y) 7→ H(x, y)
(1) H is called a bilinear form on V if H is linear in both variables, i.e., for all
xi , yj , x, y ∈ V, ai , bi ∈ F
we have
H(a1 x1 + a2 x2 , y) = a1 H(x1 , y) + a2 H(x2 , y)
H(x, b1 y1 + b2 y2 ) = b1 H(x, y1 ) + b2 H(x, y2 )

(2) A bilinear form H on V is symmetric if


H(x, y) = H(y, x) ∀x, y ∈ V

Theorem 10.1 (Representation Matrix).


Suppose that
B = (v1 , . . . , vn )
is a basis of a vector space V over a field F. Let
 
a11 · · · a1n
A = (aij ) =  ... . . . ... 
 
an1 · · · ann

be a matrix in mathbbMn (F).


We define the function
HA : V × V → F
n
X n
X n
X n
X
( xi vi , yi vj ) 7→ HA ( x i vi , y i vj )
i=1 j=1 i=1 j=1

where n n
X X
HA ( x i vi , yi vj )
i=1 j=1
n X
X n
:= aij xi yj
i=1 j=1
  
a11 · · · a1n y1
 .. . . . .
=(x1 , . . . , xn )  . . ..   .. 
 
an1 · · · ann yn
=X t AY

48
(1) Then HA is a bilinear form on V and called the bilinear form associated with A
(and relative to the basis B of V ).

(2) Conversely, every bilinear form H on V is of the form HA for some A in Mn (F). Indeed,
just set
aij = H(vi , vj ), A := (aij )
Then one can use the bilinearity of H, show that H = HA .
The matrix A is called the representation matrix of H relative to the basis of B¿

(3) HA is a symmetric bilinear form if and only if A is a symmetric matrix.

Definition 10.2 (Non-degenerate bilinear forms).


A bilinear form H on V is non-degenerate if for every y0 ∈ V , we have:

H(x, y0 ) = 0(∀x ∈ V ) ⇒ y0 = 0

A bilinear form H = HA is non-degenerate if and only if its representation matrix A is


invertible.

Definition 10.3 (Congruent matrices).


Two matrices A and B in Mn (F) are congruent if there is an invertible matrix P ∈ Mn (F)
such that
B = P t AP
Being congruent is an equivalent relation.
Consider the bilinear form
H : Fnc × Fnc → F
(X, Y ) 7→ X t AY
If we write
X = PY
with an invertible matrix P ∈ Mn (F) and introduce Y as a new coordinate system for Fnc ,
then
H(X1 , X2 ) = X1t AX2
= (P Y1 )t A(P Y2 )
= Y1t (P t AP )Y2
Thus, the bilinear form above would have a simpler form in new coordinates Y , if P t AP
(which is congruent to A) is simpler. This simplification is very useful in classfying all conics.

Theorem 10.2 (Weak version of Principle Axis Theorem).


Let A ∈ Mn (F) be a symmetric matrix. Then there is an invertible matrix P in Mn (F) such
that the matrix P t AP is diagonal:

P t AP = diag[d1 , . . . , dn ] =: D

49
i.e. A is congruent to a diagonal matrix D.
In this case, the bilinear form
n X
X n
H(X1 , X2 ) = aij xi yj
i=1 j=1

= X1t AX2
= Y1t DY2
  
d1 · · · 0 y2 1
= (y11 , . . . , y1n )  0 . . 0   ... 
.
   
0 · · · dn y2 n
Xn
= dj y1j y2j
j=1

where we have used the substitution:

Xi = P Y i

Definition 10.4 (Inner Product, Orthogonal, Norm).


We start with the real version.
Consider a function H:
V ×V →R
(x, y) 7→ H(x, y)
on a vector space V over the field R of real numbers.
The function H is called a real inner product and V a real inner product space, if the
following three conditions are satisfied, where we denote

hx, yi := H(x, y)

(1) H is a bilinear form, i.e., for all

xi , yj , x, y ∈ V, ai , bi ∈ R

we have
ha1 x1 + a2 x2 , yi = a1 hx1 , yi + a2 hx2 , yi
hx, b1 y1 + b2 y2 i = b1 hx, y1 i + b2 hx, y2 i

(2) H is symmetric, i.e., for all x, y ∈ V , we have

hx, yi = hy, xi

(3) Positivity:
For all 0 6= x ∈ V , we have
hx, xi > 0

50
Next is the complex version. Consider a function H:

V ×V →C
(x, y) 7→ H(x, y)

on a vector space V over the field C of real numbers.


The function H is called a complex inner product and V a complex inner product
space, if the following three conditions are satisfied, where we denote

hx, yi := H(x, y)

(1) H is a bilinear form, i.e., for all

xi , yj , x, y ∈ V, ai , bi ∈ R

we have
h< a1 x1 + a2 x2 , yi = a1 hx1 , yi + a2 hx2 , yi
h< x, b1 y1 + b2 y2 i = b̄1 hx, y1 i + b̄2 hx, y2 i

(2) H is symmetric, i.e., for all x, y ∈ V , we have

hx, yi = hy, xi

(3) Positivity:
For all 0 6= x ∈ V , we have
hx, xi > 0

We have three more definitions:


(1) The norm of a vector x ∈ V is denoted and defined as:
p
kxk = hx, xi

We have
kxk ≥ 0
and
kxk == 0 ⇔ x = 0V

(2) Two vectors x, y in V are orthogonal to each other and denoted as

x⊥y

if their inner product


hx, yi = 0

(3) Sometimes, we use


(V, h, i)
to denote a vector space V with an inner product h, i.

51
Definition 10.5 (Non-degenerate Inner Product).
Let (V, h, i) be an inner product space over a field F with F = R or F = C. Then the product
h, i is non-degenerate in the sense:
for every u0 ∈ V
hu0 , yi = 0(∀y ∈ V ) ⇒ u0 = 0V
and for every v0 ∈ V ,
hx, v0 i = 0(∀x ∈ V ) ⇒ v0 = 0V
Definition 10.6 (Orthonormal basis).
Let (V, h, i) be a real or complex inner product space. A basis B = (v1 , . . . , vn is called an
orthonormal basis of the inner product space V , if it satisfies the following two conditions:
(1) Orthogonality:
for all i 6= j, we have:
vi ⊥ vj i.e., hvi , vj i = 0

(2) Normalised:
for all i, we have
kvi k = 1
Namely, vi is a unit vector.
Theorem 10.3 (Gram-Schmidt Process).
Let V = Fnc with F = R or C. Employ the standard inner product h, i for V .
Let
(u1 , . . . , ur )
be a basis of a subspace W of V . Then one can apply the following Gram-Schmidt process
to get an orthonormal basis
(v1 , . . . , vr )
of W .
v10 = u1
hu2 , v10 i 0
v20 = u2 − v1
kv10 k2
k−1
X huk , v0 i
vk0 = uk − 0 2
i

i=1
kvi k
vj0
vj =
v0

j

Definition 10.7 (Adjoint matrices A∗ ).


For a matrix A = (aij ) ∈ Mn (C), the adjoint of A is defined as

A∗ = (A)t = (aij )t

i.e., the (i, j)-entry of A∗ equals aji . Note that

A∗ = (At )

52
Theorem 10.4 (Adjoint matrix A∗ and inner product).
Let V = F)nc with F = R or C. Employ the standard inner product h, i for V . For a matrix
A ∈ Mn (F), we have
hAX, Y i = hX, A∗ Y i
Theorem 10.5 (Adjoint Linear Operator).
Let T : V → V be a linear operator on an n-dimensional inner product space V over a field
F. Then we have:
(1) There is a unique linear operator

T∗ : V → V

on V such that
hT (u), vi = hu, T ∗ (v)i
Such T ∗ is called the adjoint linear operator of T .

(2) Let B = (w1 , . . . , wn ) be an orthonormal basis of the inner product space V . Then

[T ∗ ]B = ([T ]B )∗

Theorem 10.6 (Adjoint of adjoint). (T ∗ )∗ = T .


Theorem 10.7 (Adjoint of linear map combinations).

(1) Suppose that T = αIV is a scalar map. Then

T ∗ = αIV

(2)
(a1 T1 + a2 T2 )∗ = a1 T1∗ + a2 T2∗

(3)
(T1 ◦ T2 )∗ = T2∗ ◦ T1∗

Definition 10.8 (Orthogonal, Unitary, Self-adjoint, Normal linear operators).


Let A ∈ Mn (C) (resp. let T : V → V be a linear operator on an n-dimensional inner product
space over a field F = R or C and with an orthonormal basis B). Let A∗ (resp. T ∗ ) be the
adjoint of A (resp. T ).
(1) A linear operator T over a real inner product space is orthogonal if

T T ∗ = IV

(2) A real matrix A in Mn (R) is orthogonal if

AAt = In

53
(3) A linear operator T over a complex inner product space is unitary if

T T ∗ = IV

(4) A complex matrix A in mathbbMn (C) is unitary if

AA∗ = In

(5) T is self-adjoint if its adjoint T ∗ equals itself:

T = T∗

When the field F = R, a self adjoint operator is also called a symmetric operator.
(6) A complex matrix A ∈ Mn (C) is self-adjoint if the adjoint matrix of A equals itself:

A∗ = A

(7) A linear operator T over a complex inner product space is normal if

T T ∗ = T ∗T

(8) A complex matrix A ∈ Mn (C) is normal if

AA∗ = A∗ A

Orthogonal, Unitary, self-adjoint operators are normal.


Theorem 10.8. T is orthogonal, unitary, self-adjoint or normal if and only if its represen-
tation matrix A := [T ]B (relative to one hence every orthonormal basis B) is respectively
orthogonal, unitary, self-adjoint and normal.
Theorem 10.9 (Equivalent unitary matrix definition).
For a real matrix P in Mn (C), the following are equivalent, if we employ the standard inner
product on Cnc .
(1) P is unitary, i.e. P P ∗ = In .
(2) Write
P = (p1 , . . . , pn )
where the pj are the column vectors of P . Then the column vectors p1 , . . . , pn form
an orthonormal basis of Cnc .
(3) The matrix transformation
TP : Cnc → Cnc
X 7→ P X
preserves the standard inner product, i.e., for all X, Y in Cnc , we have

hP X, P Y i = hX, Y i

54
(4) The matrix transformation TP preserves the distance, i.e. for all X, Y in Cnc , we have

kP X − P Y k = kX − Y k

(5) The matrix transformation TP preserves the norm, i.e. for all X in Cnc , we have

kP Xk = kXk

(6) For one and hence every orthogonal basis

B = (v1 , . . . , vn )

of Cnc , the new basis


B 0 = BP
is again an orthonormal basis of Cnc .
Theorem 10.10 (Eigenvalues of orthogonal or unitary matrices).

(1) If a real matrix P ∈ Mn (R) is orthogonal, then every zero of pP (x) has modulus equal
to 1. In particular, the determinant

|P | = ±1

(2) If a complex matrix P ∈ Mn (C) is unitary, then every eigenvalue



λi = r1 + r2 −1

of P has modulus q
|λ| = r12 + r22 = 1
In particular, the determinant |P | ∈ C has modulus 1.
Theorem 10.11 (Eigenvalue of self-adjoint linear operators).

(1) Suppose that a real matrix A ∈ Mn (R) is symmetric. Every zero of pA (x) is a real
number.

(2) Suppose a complex matrix A ∈ Mn (C) is self-adjoint. Every zero of pA (x) is a real
number.

(3) More generally, suppose that T is a self adjoint linear operator. Every zero of pT (x) is
a real number.

(4) Suppose that T is a self-adjoint linear operator. Let vi (i = 1, 2) be two eigenvectors


corresponding to two distinct eigenvalues λi of T . We have

hv1 , v2 i = 0

55
Definition 10.9 (Positive/Negative definite linear operators).
Let A ∈ Mn (C) (resp. let V be an n-dimensional inner product space which is over a field
F = R or C).
(1) T is positive definite if T is self-adjoint and

hT (v), vi > 0

(2) T is negative definite if T is self-adjoint and

hT (v), vi < 0

Thus T is negative definite if and only if −T is positive definite.

(3) A is positive definite if A is self-adjoint and

(AX)t X = X t At X > 0

(4) A is negative definite if A is self-adjoint and

(AX)t X = X t At X < 0

Thus, A is negative definite if and only if −A is positive definite.


Theorem 10.12 (Equivalent Positive-Definite Definition).
Let
A = (aij ) ∈ Mn (R)
be a symmetric real matrix. Then A is positive definite if and only if all its principal
minors
(aij )1≤i,j≤r (1 ≤ r ≤ n)
of order r have positive determinants.
Let T be a self-adjoint linera operator on an inner product space V which is over F = R or
C and with an orthonormal basis B. Set

A := [T ]B ∈ Mn (F)

Then the following are equivalent.


(1) T is positive definite.

(2) A is positive definite.

(3) Every eigenvalue of T is positive.

(4) Every eigenvalue of A is positive.

(5) One can write A as


A = C ∗C
for some invertible complex matrix C ∈ Mn (C)

56
Theorem 10.13. Let A ∈ Mn (C). The function H on V := Cnc

H :V ×V →C
(X, Y ) 7→ hX, Y i := (AX)t Y

defines an inner product on V if and only if A is positive definite.

Theorem 10.14 (Principle Axis Theorem).

1. Let T : V → V be a linear operator on a real inner product space V of dimension n.


Then T is self-adjoint (i.e., T ∗ = T ) if and only if there is an orthonormal basis B such
that
[T ]B
is a diagonal matrix in Mn (R).

2. A real matrix A ∈ Mn (R) is self-adjoint (i.e. A∗ = A) if and only if there is an


orthogonal matrix P such that

P −1 AP = P t AP

is a diagonal matrix in Mn (R).

3. Let T : V → V be a linear operator on a complex inner product space V of dimension


n. Then T is self-adjoint (i.e., T ∗ = T ) if and only if there is an orthonormal basis B
such that
[T ]B
is a diagonal matrix in Mn (C).

4. A complex matrix A ∈ Mn (C) is self-adjoint (i.e. A∗ = A) if and only if there is an


unitary matrix U such that
U −1 AU = U ∗ AU
is a diagonal matrix in Mn (C).

Theorem 10.15 (Orthogonal Complement).


Let W be a subspace of an inner product space V . Take an orthogonal basis BW of W .

(1) One can extend BW to an orthonormal basis B = (BW , B2 ) of V .

(2) B2 is an orthonormal basis of so called orthogonal complement of W :

W ⊥ := {x ∈ V | hx, vi = 0, ∀w ∈ W }

(3)
V = W ⊕ W⊥

57
Definition 10.10 (Quadratic Form).
Let V be a vector space over a field F. A function

K:V →F

or simply K(x) is a quadratic form if there is a symmetric bilinear form

H :V ×V →F

such that
K(x) = H(x, x)

Theorem 10.16 (Principle Axis Theorem of Quadratic Form).


Let n X
n
X
f (x1 , . . . , xn ) = aij xi xj
i=1 j=1

be a quadratic form in coordinates  


x1
X =  ... 
 
xn
with
A = (aij ) ∈ Mn (R)
a symmetric matrix. Then there is an orthogonal matrix P such that f has the following
standard form
f (x1 , . . . , xn ) = λ1 y12 + · · · + λn yn2
in the new coordinates  
y1
 .. 
Y :=  .  = P −1 X
yn
where λi ∈ R are the eigenvalues of A.
This standard form is unique up to relabelling of λi yi2 .

58
11 Problems
1 Let A ∈ Mn (C) be a complex matrix of order n ≥ 9 and let

f (x) := (x − 1)2 (x − 2)3 (x − 3)4

Suppose that A is self-adjoint and f (A) = 0. Find all possible minimal polynomials
mA (x) of A.

2 Let V be a finite-dimensional inner product space and T : V → V invertible. Prove


that there exists a unitary operator U and a positive operator P on V such that
T = U ◦ P.

3 AY1314Sem2 Question 6(iii)

4 AY1415Sem2 Question 8(iv) –(vi)

5 Let V be a finite dimensional vector space over a field F and let T be a linear operator
on V . Suppose there exists v ∈ V such that {v, T (v), . . . , T n−1 (v)} is a basis for V
where n = dim(V ).

(a) Prove that the linear operators IV , T, . . . , T n−1 are linearly independent. (Done)
(b) Let S be a linear operator on V such that S ◦ T = T ◦ S. Write

S(v) = a0 v + a1 T (v) + · · · + an−1 T n−1 (v)

where a0 , a1 , · · · , an−1 ∈ F.
Prove that S = p(T ) where p(x) = a0 + a1 x + · · · + an−1 xn−1 .
(c) Suppose pT (x) = (x−λ1 )r1 (x−λ2 )r2 · · · (x−λk )rk where λ1 , λ2 , . . . , λk are distinct
eigenvalues of T . Find mT (x).

6 Let A be an invertible n × n matrix over a field F.

(a) Show that cA−1 (x) = xn [cA (0)]−1 cA ( x1 ) (Done)


(b) Show that mA−1 (x) = xk [mA (0)]−1 mA ( x1 ).

59

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy