Part IB Linear Algebra
Based on lectures by S. J. Wadsley
Notes taken by Dexter Chua
Michaelmas 2015
These notes are not endorsed by the lecturers, and I have modified them (often
significantly) after lectures. They are nowhere near accurate representations of what
was actually lectured, and in particular, all errors are almost surely mine.
Definition of a vector space (over
R
or
C
), subspaces, the space spanned by a subset.
Linear independence, bases, dimension. Direct sums and complementary subspaces.
[3]
Linear maps, isomorphisms. Relation between rank and nullity. The space of linear
maps from
U
to
V
, representation by matrices. Change of basis. Row rank and column
rank. [4]
Determinant and trace of a square matrix. Determinant of a product of two matrices
and of the inverse matrix. Determinant of an endomorphism. The adjugate matrix. [3]
Eigenvalues and eigenvectors. Diagonal and triangular forms. Characteristic and
minimal polynomials. Cayley-Hamilton Theorem over
C
. Algebraic and geometric
multiplicity of eigenvalues. Statement and illustration of Jordan normal form. [4]
Dual of a finite-dimensional vector space, dual bases and maps. Matrix representation,
rank and determinant of dual map. [2]
Bilinear forms. Matrix representation, change of basis. Symmetric forms and their link
with quadratic forms. Diagonalisation of quadratic forms. Law of inertia, classification
by rank and signature. Complex Hermitian forms. [4]
Inner product spaces, orthonormal sets, orthogonal projection,
V
=
W W
. Gram-
Schmidt orthogonalisation. Adjoints. Diagonalisation of Hermitian matrices. Orthogo-
nality of eigenvectors and properties of eigenvalues. [4]
Contents
0 Introduction
1 Vector spaces
1.1 Definitions and examples
1.2 Linear independence, bases and the Steinitz exchange lemma
1.3 Direct sums
2 Linear maps
2.1 Definitions and examples
2.2 Linear maps and matrices
2.3 The first isomorphism theorem and the rank-nullity theorem
2.4 Change of basis
2.5 Elementary matrix operations
3 Duality
3.1 Dual space
3.2 Dual maps
4 Bilinear forms I
5 Determinants of matrices
6 Endomorphisms
6.1 Invariants
6.2 The minimal polynomial
6.2.1 Aside on polynomials
6.2.2 Minimal polynomial
6.3 The Cayley-Hamilton theorem
6.4 Multiplicities of eigenvalues and Jordan normal form
7 Bilinear forms II
7.1 Symmetric bilinear forms and quadratic forms
7.2 Hermitian form
8 Inner product spaces
8.1 Definitions and basic properties
8.2 Gram-Schmidt orthogonalization
8.3 Adjoints, orthogonal and unitary maps
8.4 Spectral theory
0 Introduction
In IA Vectors and Matrices, we have learnt about vectors (and matrices) in a
rather applied way. A vector was just treated as a “list of numbers” representing
a point in space. We used these to represent lines, conics, planes and many
other geometrical notions. A matrix is treated as a “physical operation” on
vectors that stretches and rotates them. For example, we studied the properties
of rotations, reflections and shears of space. We also used matrices to express
and solve systems of linear equations. We mostly took a practical approach in
the course.
In IB Linear Algebra, we are going to study vectors in an abstract setting.
Instead of treating vectors as “lists of numbers”, we view them as things we
can add and scalar-multiply. We will write down axioms governing how these
operations should behave, just like how we wrote down the axioms of group
theory. Instead of studying matrices as an array of numbers, we instead look at
linear maps between vector spaces abstractly.
In the course, we will, of course, prove that this abstract treatment of linear
algebra is just “the same as” our previous study of “vectors as a list of numbers”.
Indeed, in certain cases, results are much more easily proved by working with
matrices (as an array of numbers) instead of abstract linear maps, and we don’t
shy away from doing so. However, most of the time, looking at these abstractly
will provide a much better fundamental understanding of how things work.
1 Vector spaces
1.1 Definitions and examples
Notation. We will use F to denote an arbitrary field, usually R or C.
Intuitively, a vector space V over a field F (or an F-vector space) is a space
with two operations:
We can add two vectors v
1
, v
2
V to obtain v
1
+ v
2
V .
We can multiply a scalar λ F with a vector v V to obtain λv V .
Of course, these two operations must satisfy certain axioms before we can
call it a vector space. However, before going into these details, we first look at a
few examples of vector spaces.
Example.
(i) R
n
=
{column vectors of length n with coefficients in R}
with the usual
addition and scalar multiplication is a vector space.
An
m × n
matrix
A
with coefficients in
R
can be viewed as a linear map
from R
m
to R
n
via v 7→ Av.
This is a motivational example for vector spaces. When confused about
definitions, we can often think what the definition means in terms of
R
n
and matrices to get some intuition.
(ii)
Let
X
be a set and define
R
X
=
{f
:
X R}
with addition (
f
+
g
)(
x
) =
f
(
x
) +
g
(
x
) and scalar multiplication (
λf
)(
x
) =
λf
(
x
). This is a vector
space.
More generally, if
V
is a vector space,
X
is a set, we can define
V
X
=
{f
:
X V } with addition and scalar multiplication as above.
(iii) Let [a, b] R be a closed interval, then
C([a, b], R) = {f R
[a,b]
: f is continuous}
is a vector space, with operations as above. We also have
C
([a, b], R) = {f R
[a,b]
: f is infinitely differentiable}
(iv)
The set of
m × n
matrices with coefficients in
R
is a vector space, using
componentwise addition and scalar multiplication, is a vector space.
Of course, we cannot take a random set, define some random operations
called addition and scalar multiplication, and call it a vector space. These
operations have to behave sensibly.
Definition
(Vector space)
.
An
F
-vector space is an (additive) abelian group
V
together with a function F × V V , written (λ, v) 7→ λv , such that
(i) λ(µv) = λµv for all λ, µ F, v V (associativity)
(ii) λ(u + v) = λu + λv for all λ F, u, v V (distributivity in V )
(iii) (λ + µ)v = λv + µv for all λ, µ F, v V (distributivity in F)
(iv) 1v = v for all v V (identity)
We always write
0
for the additive identity in
V
, and call this the identity.
By abuse of notation, we also write 0 for the trivial vector space {0}.
In a general vector space, there is no notion of “coordinates”, length, angle
or distance. For example, it would be difficult to assign these quantities to the
vector space of real-valued continuous functions in [a, b].
From the axioms, there are a few results we can immediately prove.
Proposition.
In any vector space
V
, 0
v
=
0
for all
v V
, and (
1)
v
=
v
,
where v is the additive inverse of v.
Proof is left as an exercise.
In mathematics, whenever we define “something”, we would also like to define
a “sub-something”. In the case of vector spaces, this is a subspace.
Definition
(Subspace)
.
If
V
is an
F
-vector space, then
U V
is an (
F
-linear)
subspace if
(i) u, v U implies u + v U .
(ii) u U, λ F implies λu U .
(iii) 0 U .
These conditions can be expressed more concisely as
U
is non-empty and if
λ, µ F, u, v U , then λu + µv U”.
Alternatively,
U
is a subspace of
V
if it is itself a vector space, inheriting the
operations from V .
We sometimes write U V if U is a subspace of V .
Example.
(i) {(x
1
, x
2
, x
3
) R
3
: x
1
+ x
2
+ x
3
= t} is a subspace of R
3
iff t = 0.
(ii)
Let
X
be a set. We define the support of
f
in
F
X
to be
supp
(
f
) =
{x X
:
f
(
x
)
6
= 0
}
. Then the set of functions with finite support
forms a vector subspace. This is since
supp
(
f
+
g
)
supp
(
f
)
supp
(
g
),
supp(λf) = supp(f) (for λ 6= 0) and supp(0) = .
If we have two subspaces
U
and
V
, there are several things we can do with
them. For example, we can take the intersection
U V
. We will shortly show
that this will be a subspace. However, taking the union will in general not
produce a vector space. Instead, we need the sum:
Definition
(Sum of subspaces)
.
Suppose
U, W
are subspaces of an
F
vector
space V . The sum of U and V is
U + W = {u + w : u U, w W }.
Proposition.
Let
U, W
be subspaces of
V
. Then
U
+
W
and
U W
are
subspaces.
Proof. Let u
i
+ w
i
U + W , λ, µ F. Then
λ(u
1
+ w
1
) + µ(u
2
+ w
2
) = (λu
1
+ µu
2
) + (λw
1
+ µw
2
) U + W.
Similarly, if
v
i
U W
, then
λv
1
+
µv
2
U
and
λv
1
+
µv
2
W
. So
λv
1
+ µv
2
U W .
Both U W and U + W contain 0, and are non-empty. So done.
In addition to sub-somethings, we often have quotient-somethings as well.
Definition
(Quotient spaces)
.
Let
V
be a vector space, and
U V
a subspace.
Then the quotient group
V/U
can be made into a vector space called the quotient
space, where scalar multiplication is given by (λ, v + U ) = (λv) + U.
This is well defined since if
v
+
U
=
w
+
U V /U
, then
v w U
. Hence
for λ F, we have λv λw U. So λv + U = λw + U .
1.2 Linear independence, bases and the Steinitz exchange
lemma
Recall that in
R
n
, we had the “standard basis” made of vectors of the form
e
i
= (0
, ··· ,
0
,
1
,
0
, ··· ,
0), with 1 in the
i
th component and 0 otherwise. We
call this a basis because everything in
R
n
can be (uniquely) written as a sum
of (scalar multiples of) these basis elements. In other words, the whole
R
n
is
generated by taking sums and multiples of the basis elements.
We would like to capture this idea in general vector spaces. The most
important result in this section is to prove that for any vector space
V
, any two
basis must contain the same number of elements. This means we can define the
“dimension” of a vector space as the number of elements in the basis.
While this result sounds rather trivial, it is a very important result. We will
in fact prove a slightly stronger statement than what was stated above, and this
ensures that the dimension of a vector space is well-behaved. For example, the
subspace of a vector space has a smaller dimension than the larger space (at
least when the dimensions are finite).
This is not the case when we study modules in IB Groups, Rings and Modules,
which are generalizations of vector spaces. Not all modules have basis, which
makes it difficult to define the dimension. Even for those that have basis, the
behaviour of the “dimension” is complicated when, say, we take submodules.
The existence and well-behavedness of basis and dimension is what makes linear
algebra different from modules.
Definition
(Span)
.
Let
V
be a vector space over
F
and
S V
. The span of
S
is defined as
hSi =
(
n
X
i=1
λ
i
s
i
: λ
i
F, s
i
S, n 0
)
This is the smallest subspace of V containing S.
Note that the sums must be finite. We will not play with infinite sums, since
the notion of convergence is not even well defined in a general vector space.
Example.
(i) Let V = R
3
and S =
1
0
0
,
0
1
1
,
1
2
2
. Then
hSi =
a
b
b
: a, b R
.
Note that any subset of S of order 2 has the same span as S.
(ii) Let X be a set, x X. Define the function δx : X F by
δx(y) =
(
1 y = x
0 y 6= x
.
Then hδx : x Xi is the set of all functions with finite support.
Definition
(Spanning set)
.
Let
V
be a vector space over
F
and
S V
.
S
spans
V if hSi = V .
Definition
(Linear independence)
.
Let
V
be a vector space over
F
and
S V
.
Then S is linearly independent if whenever
n
X
i=1
λ
i
s
i
= 0 with λ
i
F, s
1
, s
2
, ··· , s
n
S distinct,
we must have λ
i
= 0 for all i.
If S is not linearly independent, we say it is linearly dependent.
Definition
(Basis)
.
Let
V
be a vector space over
F
and
S V
. Then
S
is a
basis for V if S is linearly independent and spans V .
Definition
(Finite dimensional)
.
A vector space is finite dimensional if there
is a finite basis.
Ideally, we would want to define the dimension as the number of vectors in
the basis. However, we must first show that this is well-defined. It is certainly
plausible that a vector space has a basis of size 7 as well as a basis of size 3. We
must show that this can never happen, which is something we’ll do soon.
We will first have an example:
Example.
Again, let
V
=
R
3
and
S
=
1
0
0
,
0
1
1
,
1
2
2
. Then
S
is
linearly dependent since
1
1
0
0
+ 2
0
1
1
+ (1)
1
2
2
= 0.
S also does not span V since
0
0
1
6∈ hSi.
Note that no linearly independent set can contain
0
, as 1
· 0
=
0
. We also
have h∅i = {0} and is a basis for this space.
There is an alternative way in which we can define linear independence.
Lemma. S V
is linearly dependent if and only if there are distinct
s
0
, ··· , s
n
S and λ
1
, ··· , λ
n
F such that
n
X
i=1
λ
i
s
i
= s
0
.
Proof.
If
S
is linearly dependent, then there are some
λ
1
, ··· , λ
n
F
not all
zero and s
1
, ··· , s
n
S such that
P
λ
i
s
i
= 0. Wlog, let λ
1
6= 0. Then
s
1
=
n
X
i=2
λ
i
λ
1
s
i
.
Conversely, if s
0
=
P
n
i=1
λ
i
s
i
, then
(1)s
0
+
n
X
i=1
λ
i
s
i
= 0.
So S is linearly dependent.
This in turn gives an alternative characterization of what it means to be a
basis:
Proposition.
If
S
=
{e
1
, ··· , e
n
}
is a subset of
V
over
F
, then it is a basis if
and only if every
v V
can be written uniquely as a finite linear combination
of elements in S, i.e. as
v =
n
X
i=1
λ
i
e
i
.
Proof.
We can view this as a combination of two statements: it can be spanned
in at least one way, and it can be spanned in at most one way. We will see that
the first part corresponds to
S
spanning
V
, and the second part corresponds to
S being linearly independent.
In fact,
S
spanning
V
is defined exactly to mean that every item
v V
can
be written as a finite linear combination in at least one way.
Now suppose that S is linearly independent, and we have
v =
n
X
i=1
λ
i
e
i
=
n
X
i=1
µ
i
e
i
.
Then we have
0 = v v =
n
X
i=1
(λ
i
µ
i
)e
i
.
Linear independence implies that
λ
i
µ
i
= 0 for all
i
. Hence
λ
i
=
µ
i
. So
v
can
be expressed in a unique way.
On the other hand, if S is not linearly independent, then we have
0 =
n
X
i=1
λ
i
e
i
where λ
i
6= 0 for some i. But we also know that
0 =
n
X
i=1
0 · e
i
.
So there are two ways to write 0 as a linear combination. So done.
Now we come to the key theorem:
Theorem
(Steinitz exchange lemma)
.
Let
V
be a vector space over
F
, and
S
=
{e
1
, ··· , e
n
}
a finite linearly independent subset of
V
, and
T
a spanning
subset of
V
. Then there is some
T
0
T
of order
n
such that (
T \ T
0
)
S
still
spans V . In particular, |T | n.
What does this actually say? This says if
T
is spanning and
S
is independent,
there is a way of grabbing
|S|
many elements away from
T
and replace them
with S, and the result will still be spanning.
In some sense, the final remark is the most important part. It tells us that
we cannot have a independent set larger than a spanning set, and most of our
corollaries later will only use this remark.
This is sometimes stated in the following alternative way for |T | < .
Corollary.
Let
{e
1
, ··· , e
n
}
be a linearly independent subset of
V
, and sup-
pose
{f
1
, ··· , f
m
}
spans
V
. Then there is a re-ordering of the
{f
i
}
such that
{e
1
, ··· , e
n
, f
n+1
, ··· , f
m
} spans V .
The proof is going to be slightly technical and notationally daunting. So it
helps to give a brief overview of what we are going to do in words first. The idea
is to do the replacement one by one. The first one is easy. Start with
e
1
. Since
T is spanning, we can write
e
1
=
X
λ
i
t
i
for some
t
i
T, λ
i
F
non-zero. We then replace with
t
1
with
e
1
. The result is
still spanning, since the above formula allows us to write
t
1
in terms of
e
1
and
the other t
i
.
We continue inductively. For the rth element, we again write
e
r
=
X
λ
i
t
i
.
We would like to just pick a random
t
i
and replace it with
e
r
. However, we
cannot do this arbitrarily, since the lemma wants us to replace something in T
with with
e
r
. After all that replacement procedure before, some of the
t
i
might
have actually come from S.
This is where the linear independence of
S
kicks in. While some of the
t
i
might be from S, we cannot possibly have all all of them being from S, or else
this violates the linear independence of
S
. Hence there is something genuinely
from T , and we can safely replace it with e
r
.
We now write this argument properly and formally.
Proof.
Suppose that we have already found
T
0
r
T
of order 0
r < n
such that
T
r
= (T \ T
0
r
) {e
1
, ··· , e
r
}
spans V .
(Note that the case
r
= 0 is trivial, since we can take
T
0
r
=
, and the case
r = n is the theorem which we want to achieve.)
Suppose we have these. Since T
r
spans V , we can write
e
r+1
=
k
X
i=1
λ
i
t
i
, λ
i
F, t
i
T
r
.
We know that the
e
i
are linearly independent, so not all
t
i
’s are
e
i
’s. So there is
some j such that t
j
(T \ T
0
r
). We can write this as
t
j
=
1
λ
j
e
r+1
+
X
i6=j
λ
i
λ
j
t
i
.
We let T
0
r+1
= T
0
r
{t
j
} of order r + 1, and
T
r+1
= (T \ T
0
r+1
) {e
1
, ··· , e
r+1
} = (T
r
\ {t
j
}} {e
r+1
}
Since t
j
is in the span of T
r
{e
r+1
}, we have t
j
hT
r+1
i. So
V hT
r+1
i hT
r
i = V.
So hT
r+1
i = V .
Hence we can inductively find T
n
.
From this lemma, we can immediately deduce a lot of important corollaries.
Corollary. Suppose V is a vector space over F with a basis of order n. Then
(i) Every basis of V has order n.
(ii) Any linearly independent set of order n is a basis.
(iii) Every spanning set of order n is a basis.
(iv) Every finite spanning set contains a basis.
(v) Every linearly independent subset of V can be extended to basis.
Proof. Let S = {e
1
, ··· , e
n
} be the basis for V .
(i)
Suppose
T
is another basis. Since
S
is independent and
T
is spanning,
|T | |S|.
The other direction is less trivial, since
T
might be infinite, and Steinitz
does not immediately apply. Instead, we argue as follows: since
T
is
linearly independent, every finite subset of
T
is independent. Also,
S
is
spanning. So every finite subset of
T
has order at most
|S|
. So
|T | |S|
.
So |T | = |S|.
(ii)
Suppose now that
T
is a linearly independent subset of order
n
, but
hT i 6
=
V
. Then there is some
v V \ hT i
. We now show that
T {v}
is
independent. Indeed, if
λ
0
v +
m
X
i=1
λ
i
t
i
= 0
with λ
i
F, t
1
, ··· , t
m
T distinct, then
λ
0
v =
m
X
i=1
(λ
i
)t
i
.
Then
λ
0
v hT i
. So
λ
0
= 0. As
T
is linearly independent, we have
λ
0
=
···
=
λ
m
= 0. So
T {v}
is a linearly independent subset of size
> n. This is a contradiction since S is a spanning set of size n.
(iii)
Let
T
be a spanning set of order
n
. If
T
were linearly dependent, then
there is some t
0
, ··· , t
m
T distinct and λ
1
, ··· , λ
m
F such that
t
0
=
X
λ
i
t
i
.
So
t
0
hT \ {t
0
}i
, i.e.
hT \ {t
0
}i
=
V
. So
T \ {t
0
}
is a spanning set of
order n 1, which is a contradiction.
(iv)
Suppose
T
is any finite spanning set. Let
T
0
T
be a spanning set of least
possible size. This exists because
T
is finite. If
|T
0
|
has size
n
, then done
by (iii). Otherwise by the Steinitz exchange lemma, it has size
|T
0
| > n
.
So
T
0
must be linearly dependent because
S
is spanning. So there is some
t
0
, ··· , t
m
T
distinct and
λ
1
, ··· , λ
m
F
such that
t
0
=
P
λ
i
t
i
. Then
T
0
\ {t
0
} is a smaller spanning set. Contradiction.
(v)
Suppose
T
is a linearly independent set. Since
S
spans, there is some
S
0
S
of order
|T |
such that (
S \S
0
)
T
spans
V
by the Steinitz exchange
lemma. So by (ii), (S \ S
0
) T is a basis of V containing T .
Note that the last part is where we actually use the full result of Steinitz.
Finally, we can use this to define the dimension.
Definition
(Dimension)
.
If
V
is a vector space over
F
with finite basis
S
, then
the dimension of V , written
dim V = dim
F
V = |S|.
By the corollary,
dim V
does not depend on the choice of
S
. However, it does
depend on
F
. For example,
dim
C
C
= 1 (since
{
1
}
is a basis), but
dim
R
C
= 2
(since {1, i} is a basis).
After defining the dimension, we can prove a few things about dimensions.
Lemma.
If
V
is a finite dimensional vector space over
F
,
U V
is a proper
subspace, then U is finite dimensional and dim U < dim V .
Proof.
Every linearly independent subset of
V
has size at most
dim V
. So let
S U
be a linearly independent subset of largest size. We want to show that
S
spans U and |S| < dim V .
If
v V \hSi
, then
S {v}
is linearly independent. So
v 6∈ U
by maximality
of S. This means that hSi = U.
Since
U 6
=
V
, there is some
v V \ U
=
V \ hSi
. So
S {v}
is a linearly
independent subset of order
|S|
+ 1. So
|S|
+ 1
dim V
. In particular,
dim U
=
|S| < dim V .
Proposition.
If
U, W
are subspaces of a finite dimensional vector space
V
, then
dim(U + W ) = dim U + dim W dim(U W ).
The proof is not hard, as long as we manage to pick the right basis to do the
proof. This is our slogan:
When you choose a basis, always choose the right basis.
We need a basis for all four of them, and we want to compare the bases. So we
want to pick bases that are compatible.
Proof.
Let
R
=
{v
1
, ··· , v
r
}
be a basis for
U W
. This is a linearly independent
subset of U . So we can extend it to be a basis of U by
S = {v
1
, ··· , v
r
, u
r+1
, ··· , u
s
}.
Similarly, for W , we can obtain a basis
T = {v
1
, ··· , v
r
, w
r+1
, ··· , w
t
}.
We want to show that
dim
(
U
+
W
) =
s
+
t r
. It is sufficient to prove that
S T is a basis for U + W .
We first show spanning. Suppose
u
+
w U
+
W
,
u U, w W
. Then
u hSi and w hT i. So u + w hS T i. So U + W = hS T i.
To show linear independence, suppose we have a linear relation
r
X
i=1
λ
i
v
i
+
s
X
j=r+1
µ
j
u
j
+
t
X
k=r+1
ν
k
w
k
= 0.
So
X
λ
i
v
i
+
X
µ
j
u
j
=
X
ν
k
w
k
.
Since the left hand side is something in
U
, and the right hand side is something
in W , they both lie in U W .
Since
S
is a basis of
U
, there is only one way of writing the left hand vector
as a sum of
v
i
and
u
j
. However, since
R
is a basis of
U W
, we can write the
left hand vector just as a sum of
v
i
’s. So we must have
µ
j
= 0 for all
j
. Then
we have
X
λ
i
v
i
+
X
ν
k
w
k
= 0.
Finally, since
T
is linearly independent,
λ
i
=
ν
k
= 0 for all
i, k
. So
S T
is
linearly independent.
Proposition.
If
V
is a finite dimensional vector space over
F
and
U V
is a
subspace, then
dim V = dim U + dim V/U.
We can view this as a linear algebra version of Lagrange’s theorem. Combined
with the first isomorphism theorem for vector spaces, this gives the rank-nullity
theorem.
Proof.
Let
{u
1
, ··· , u
m
}
be a basis for
U
and extend this to a basis
{u
1
, ··· , u
m
,
v
m+1
, ··· , v
n
}
for
V
. We want to show that
{v
m+1
+
U, ··· , v
n
+
U}
is a basis
for V /U.
It is easy to see that this spans V /U. If v + U V/U, then we can write
v =
X
λ
i
u
i
+
X
µ
i
v
i
.
Then
v + U =
X
µ
i
(v
i
+ U ) +
X
λ
i
(u
i
+ U ) =
X
µ
i
(v
i
+ U ).
So done.
To show that they are linearly independent, suppose that
X
λ
i
(v
i
+ U ) = 0 + U = U.
Then this requires
X
λ
i
v
i
U.
Then we can write this as a linear combination of the u
i
’s. So
X
λ
i
v
i
=
X
µ
j
u
j
for some
µ
j
. Since
{u
1
, ··· , u
m
, v
n+1
, ··· , v
n
}
is a basis for
V
, we must have
λ
i
= µ
j
= 0 for all i, j. So {v
i
+ U } is linearly independent.
1.3 Direct sums
We are going to define direct sums in many ways in order to confuse students.
Definition
((Internal) direct sum)
.
Suppose
V
is a vector space over
F
and
U, W V
are subspaces. We say that
V
is the (internal) direct sum of
U
and
W if
(i) U + W = V
(ii) U W = 0.
We write V = U W .
Equivalently, this requires that every
v V
can be written uniquely as
u
+
w
with u U, w W . We say that U and W are complementary subspaces of V .
You will show in the example sheets that given any subspace
U V
,
U
must
have a complementary subspace in V .
Example.
Let
V
=
R
2
, and
U
=
h
0
1
i
. Then
h
1
1
i
and
h
1
0
i
are both
complementary subspaces to U in V .
Definition
((External) direct sum)
.
If
U, W
are vector spaces over
F
, the
(external) direct sum is
U W = {(u, w) : u U, w W },
with addition and scalar multiplication componentwise:
(u
1
, w
1
) + (u
2
, w
2
) = (u
1
+ u
2
, w
1
+ w
2
), λ(u, w) = (λu, λw).
The difference between these two definitions is that the first is decomposing
V
into smaller spaces, while the second is building a bigger space based on two
spaces.
Note, however, that the external direct sum
U W
is the internal direct
sum of
U
and
W
viewed as subspaces of
U W
, i.e. as the internal direct sum
of
{
(
u, 0
) :
u U}
and
{
(
0, v
) :
v V }
. So these two are indeed compatible
notions, and this is why we give them the same name and notation.
Definition
((Multiple) (internal) direct sum)
.
If
U
1
, ··· , U
n
V
are subspaces
of V , then V is the (internal) direct sum
V = U
1
··· U
n
=
n
M
i=1
U
i
if every v V can be written uniquely as v =
P
u
i
with u
i
U
i
.
This can be extended to an infinite sum with the same definition, just noting
that the sum v =
P
u
i
has to be finite.
For more details, see example sheet 1 Q. 10, where we prove in particular
that dim V =
P
dim U
i
.
Definition
((Multiple) (external) direct sum)
.
If
U
1
, ··· , U
n
are vector spaces
over F, the external direct sum is
U
1
··· U
n
=
n
M
i=1
U
i
= {(u
1
, ··· , u
n
) : u
i
U
i
},
with pointwise operations.
This can be made into an infinite sum if we require that all but finitely many
of the u
i
have to be zero.
2 Linear maps
In mathematics, apart from studying objects, we would like to study functions
between objects as well. In particular, we would like to study functions that
respect the structure of the objects. With vector spaces, the kinds of functions
we are interested in are linear maps.
2.1 Definitions and examples
Definition
(Linear map)
.
Let
U, V
be vector spaces over
F
. Then
α
:
U V
is a linear map if
(i) α(u
1
+ u
2
) = α(u
1
) + α(u
2
) for all u
i
U .
(ii) α(λu) = λα(u) for all λ F, u U.
We write L(U, V ) for the set of linear maps U V .
There are a few things we should take note of:
If we are lazy, we can combine the two requirements to the single require-
ment that
α(λu
1
+ µu
2
) = λα(u
1
) + µα(u
2
).
It is easy to see that if
α
is linear, then it is a group homomorphism (if we
view vector spaces as groups). In particular, α(0) = 0.
If we want to stress the field
F
, we say that
α
is
F
-linear. For example,
complex conjugation is a map C C that is R-linear but not C -linear.
Example.
(i)
Let
A
be an
n×m
matrix with coefficients in
F
. We will write
A M
n,m
(
F
).
Then α : F
m
F
n
defined by v Av is linear.
Recall matrix multiplication is defined by: if A
ij
is the ijth coefficient of
A, then the ith coefficient of Av is A
ij
v
j
. So we have
α(λu + µv)
i
=
m
X
j=1
A
ij
(λu + µv)
j
= λ
m
X
j=1
A
ij
u
j
+ µ
m
X
j=1
A
ij
v
j
= λα(u)
i
+ µα(v)
i
.
So α is linear.
(ii)
Let
X
be a set and
g F
X
. Then we define
m
g
:
F
X
F
X
by
m
g
(
f
)(
x
) =
g(x)f(x). Then m
g
is linear. For example, f (x) 7→ 2x
2
f(x) is linear.
(iii)
Integration
I
: (
C
([
a, b
])
, R
)
(
C
([
a, b
])
, R
) defined by
f 7→
R
x
a
f
(
t
) d
t
is
linear.
(iv) Differentiation D : (C
([a, b]), R) (C
([a, b]), R) by f 7→ f
0
is linear.
(v)
If
α, β L
(
U, V
), then
α
+
β
defined by (
α
+
β
)(
u
) =
α
(
u
) +
β
(
u
) is linear.
Also, if λ F, then λα defined by (λα)(u) = λ(α(u)) is also linear.
In this way, L(U, V ) is also a vector space over F.
(vi)
Composition of linear maps is linear. Using this, we can show that many
things are linear, like differentiating twice, or adding and then multiplying
linear maps.
Just like everything else, we want to define isomorphisms.
Definition
(Isomorphism)
.
We say a linear map
α
:
U V
is an isomorphism
if there is some β : V U (also linear) such that α β = id
V
and β α = id
U
.
If there exists an isomorphism
U V
, we say
U
and
V
are isomorphic, and
write U
=
V .
Lemma.
If
U
and
V
are vector spaces over
F
and
α
:
U V
, then
α
is an
isomorphism iff α is a bijective linear map.
Proof.
If
α
is an isomorphism, then it is clearly bijective since it has an inverse
function.
Suppose
α
is a linear bijection. Then as a function, it has an inverse
β
:
V U
. We want to show that this is linear. Let
v
1
, v
2
V
,
λ, µ F
. We
have
αβ(λv
1
+ µv
2
) = λv
1
+ µv
2
= λαβ(v
1
) + µαβ(v
2
) = α(λβ(v
1
) + µβ(v
2
)).
Since α is injective, we have
β(λv
1
+ µv
2
) = λβ(v
1
) + µβ(v
2
).
So β is linear.
Definition
(Image and kernel)
.
Let
α
:
U V
be a linear map. Then the
image of α is
im α = {α(u) : u U }.
The kernel of α is
ker α = {u : α(u) = 0}.
It is easy to show that these are subspaces of V and U respectively.
Example.
(i)
Let
A M
m,n
(
F
) and
α
:
F
n
F
m
be the linear map
v 7→ Av
. Then the
system of linear equations
m
X
j=1
A
ij
x
j
= b
i
, 1 i n
has a solution iff (b
1
, ··· , b
n
) im α.
The kernel of α contains all solutions to
P
j
A
ij
x
j
= 0.
(ii) Let β : C
(R, R) C
(R, R) that sends
β(f )(t) = f
00
(t) + p(t)f
0
(t) + q(t)f (t).
for some p, q C
(R, R).
Then if
y
(
t
)
im β
, then there is a solution (in
C
(
R, R
)) to the differential
equation
f
00
(t) + p(t)f
0
(t) + q(t)f (t) = y(t).
Similarly, ker β contains the solutions to the homogeneous equation
f
00
(t) + p(t)f
0
(t) + q(t)f (t) = 0.
If two vector spaces are isomorphic, then it is not too surprising that they
have the same dimension, since isomorphic spaces are “the same”. Indeed this is
what we are going to show.
Proposition. Let α : U V be an F-linear map. Then
(i)
If
α
is injective and
S U
is linearly independent, then
α
(
S
) is linearly
independent in V .
(ii) If α is surjective and S U spans U, then α(S) spans V .
(iii) If α is an isomorphism and S U is a basis, then α(S) is a basis for V .
Here (iii) immediately shows that two isomorphic spaces have the same
dimension.
Proof.
(i)
We prove the contrapositive. Suppose that
α
is injective and
α
(
S
) is linearly
dependent. So there are
s
0
, ··· , s
n
S
distinct and
λ
1
, ··· , λ
n
F
not all
zero such that
α(s
0
) =
n
X
i=1
λ
i
α(s
i
) = α
n
X
i=1
λ
i
s
i
!
.
Since α is injective, we must have
s
0
=
n
X
i=1
λ
i
s
i
.
This is a non-trivial relation of the s
i
in U . So S is linearly dependent.
(ii)
Suppose
α
is surjective and
S
spans
U
. Pick
v V
. Then there is some
u U
such that
α
(
u
) =
v
. Since
S
spans
U
, there is some
s
1
, ··· , s
n
S
and λ
1
, ··· , λ
n
F such that
u =
n
X
I=1
λ
i
s
i
.
Then
v = α(u) =
n
X
i=1
λ
i
α(s
i
).
So α(S) spans V .
(iii) Follows immediately from (i) and (ii).
Corollary.
If
U
and
V
are finite-dimensional vector spaces over
F
and
α
:
U V
is an isomorphism, then dim U = dim V .
Note that we restrict it to finite-dimensional spaces since we’ve only shown
that dimensions are well-defined for finite dimensional spaces. Otherwise, the
proof works just fine for infinite dimensional spaces.
Proof.
Let
S
be a basis for
U
. Then
α
(
S
) is a basis for
V
. Since
α
is injective,
|S| = |α(S)|. So done.
How about the other way round? If two vector spaces have the same di-
mension, are they necessarily isomorphic? The answer is yes, at least for
finite-dimensional ones.
However, we will not just prove that they are isomorphic. We will show that
they are isomorphic in many ways.
Proposition.
Suppose
V
is a
F
-vector space of dimension
n <
. Then writing
e
1
, ··· , e
n
for the standard basis of F
n
, there is a bijection
Φ : {isomorphisms F
n
V } {(ordered) basis(v
1
, ··· , v
n
) for V },
defined by
α 7→ (α(e
1
), ··· , α(e
n
)).
Proof.
We first make sure this is indeed a function if
α
is an isomorphism,
then from our previous proposition, we know that it sends a basis to a basis. So
(α(e
1
), ··· , α(e
n
)) is indeed a basis for V .
We now have to prove surjectivity and injectivity.
Suppose
α, β
:
F
n
V
are isomorphism such that Φ(
α
) = Φ(
β
). In other
words, α(e
i
) = β(e
i
) for all i. We want to show that α = β. We have
α
x
1
.
.
.
x
n
= α
n
X
i=1
x
i
e
i
!
=
X
x
i
α(e
i
) =
X
x
i
β(e
i
) = β
x
1
.
.
.
x
n
.
Hence α = β.
Next, suppose that (v
1
, ··· , v
n
) is an ordered basis for V . Then define
α
x
1
.
.
.
x
n
=
X
x
i
v
i
.
It is easy to check that this is well-defined and linear. We also know that
α
is
injective since (
v
1
, ··· , v
n
) is linearly independent. So if
P
x
i
v
i
=
P
y
i
v
i
, then
x
i
=
y
i
. Also,
α
is surjective since (
v
1
, ··· , v
n
) spans
V
. So
α
is an isomorphism,
and by construction Φ(α) = (v
1
, ··· , v
n
).
2.2 Linear maps and matrices
Recall that our first example of linear maps is matrices acting on
F
n
. We will
show that in fact, all linear maps come from matrices. Since we know that all
vector spaces are isomorphic to
F
n
, this means we can represent arbitrary linear
maps on vector spaces by matrices.
This is a useful result, since it is sometimes easier to argue about matrices
than linear maps.
Proposition.
Suppose
U, V
are vector spaces over
F
and
S
=
{e
1
, ··· , e
n
}
is a
basis for
U
. Then every function
f
:
S V
extends uniquely to a linear map
U V .
The slogan is “to define a linear map, it suffices to define its values on a
basis”.
Proof.
For uniqueness, first suppose
α, β
:
U V
are linear and extend
f
:
S
V . We have sort-of proved this already just now.
If u U, we can write u =
P
n
i=1
u
i
e
i
with u
i
F since S spans. Then
α(u) = α
X
u
i
e
i
=
X
u
i
α(e
i
) =
X
u
i
f(e
i
).
Similarly,
β(u) =
X
u
i
f(e
i
).
So α(u) = β(u) for every u. So α = β.
For existence, if
u U
, we can write
u
=
P
u
i
e
i
in a unique way. So defining
α(u) =
X
u
i
f(e
i
)
is unambiguous. To show linearity, let λ, µ F, u, v U. Then
α(λu + µv) = α
X
(λu
i
+ µv
i
)e
i
=
X
(λu
i
+ µv
i
)f(e
i
)
= λ
X
u
i
f(e
i
)
+ µ
X
v
i
f(e
i
)
= λα(u) + µα(v).
Moreover, α does extend f.
Corollary.
If
U
and
V
are finite-dimensional vector spaces over
F
with bases
(e
1
, ··· , e
m
) and (f
1
, ··· , f
n
) respectively, then there is a bijection
Mat
n,m
(F) L(U, V ),
sending A to the unique linear map α such that α(e
i
) =
P
a
ji
f
j
.
We can interpret this as follows: the
i
th column of
A
tells us how to write
α(e
i
) in terms of the f
j
.
We can also draw a fancy diagram to display this result. Given a basis
e
1
, ··· , e
m
, by our bijection, we get an isomorphism
s
(
e
i
) :
U F
m
. Similarly,
we get an isomorphism s(f
i
) : V F
n
.
Since a matrix is a linear map
A
:
F
m
F
n
, given a matrix
A
, we can
produce a linear map α : U V via the following composition
U F
m
F
n
V.
s(e
i
)
A
s(f
i
)
1
We can put this into a square:
F
m
F
n
U V
A
s(e
i
)
α
s(f
i
)
Then the corollary tells us that every
A
gives rise to an
α
, and every
α
corresponds
to an A that fit into this diagram.
Proof.
If
α
is a linear map
U V
, then for each 1
i m
, we can write
α
(
e
i
)
uniquely as
α(e
i
) =
n
X
j=1
a
ji
f
j
for some
a
ji
F
. This gives a matrix
A
= (
a
ij
). The previous proposition tells
us that every matrix A arises in this way, and α is determined by A.
Definition
(Matrix representation)
.
We call the matrix corresponding to a
linear map
α L
(
U, V
) under the corollary the matrix representing
α
with
respect to the bases (e
1
, ··· , e
m
) and (f
1
, ··· , f
n
).
It is an exercise to show that the bijection
Mat
n,m
(
F
)
L
(
U, V
) is an iso-
morphism of the vector spaces and deduce that
dim L
(
U, V
) = (
dim U
)(
dim V
).
Proposition.
Suppose
U, V, W
are finite-dimensional vector spaces over
F
with
bases R = (u
1
, ··· , u
r
) , S = (v
1
, .., v
s
) and T = (w
1
, ··· , w
t
) respectively.
If
α
:
U V
and
β
:
V W
are linear maps represented by
A
and
B
respectively (with respect to
R
,
S
and
T
), then
βα
is linear and represented by
BA with respect to R and T .
F
r
F
s
F
t
U V W
A B
s(R)
α
s(S)
β
s(T )
Proof.
Verifying
βα
is linear is straightforward. Next we write
βα
(
u
i
) as a linear
combination of w
1
, ··· , w
t
:
βα(u
i
) = β
X
k
A
ki
v
k
!
=
X
k
A
ki
β(v
k
)
=
X
k
A
ki
X
j
B
jk
w
j
=
X
j
X
k
B
jk
A
ki
!
w
j
=
X
j
(BA)
ji
w
j
2.3 The first isomorphism theorem and the rank-nullity
theorem
The main theorem of this section is the rank-nullity theorem, which relates the
dimensions of the kernel and image of a linear map. This is in fact an easy
corollary of a stronger result, known as the first isomorphism theorem, which
directly relates the kernel and image themselves. This first isomorphism is an
exact analogy of that for groups, and should not be unfamiliar. We will also
provide another proof that does not involve quotients.
Theorem
(First isomorphism theorem)
.
Let
α
:
U V
be a linear map. Then
ker α
and
im α
are subspaces of
U
and
V
respectively. Moreover,
α
induces an
isomorphism
¯α : U/ ker α im α
(u + ker α) 7→ α(u)
Note that if we view a vector space as an abelian group, then this is exactly
the first isomorphism theorem of groups.
Proof. We know that 0 ker α and 0 im α.
Suppose u
1
, u
2
ker α and λ
1
, λ
2
F. Then
α(λ
1
u
1
+ λ
2
u
2
) = λ
1
α(u
1
) + λ
2
α(u
2
) = 0.
So λ
1
u
1
+ λ
2
u
2
ker α. So ker α is a subspace.
Similarly, if
α
(
u
1
)
, α
(
u
2
)
im α
, then
λα
(
u
1
)+
λ
2
α
(
u
2
) =
α
(
λ
1
u
1
+
λ
2
u
2
)
im α. So im α is a subspace.
Now by the first isomorphism theorem of groups,
¯α
is a well-defined isomor-
phism of groups. So it remains to show that
¯α
is a linear map. Indeed, we
have
¯α(λ(u + ker α)) = α(λu) = λα(u) = λ(¯α(u + ker α)).
So ¯α is a linear map.
Definition
(Rank and nullity)
.
If
α
:
U V
is a linear map between finite-
dimensional vector spaces over
F
(in fact we just need
U
to be finite-dimensional),
the rank of
α
is the number
r
(
α
) =
dim im α
. The nullity of
α
is the number
n(α) = dim ker α.
Corollary
(Rank-nullity theorem)
.
If
α
:
U V
is a linear map and
U
is
finite-dimensional, then
r(α) + n(α) = dim U.
Proof.
By the first isomorphism theorem, we know that
U/ ker α
=
im α
. So we
have
dim im α = dim(U/ ker α) = dim U dim ker α.
So the result follows.
We can also prove this result without the first isomorphism theorem, and
say a bit more in the meantime.
Proposition.
If
α
:
U V
is a linear map between finite-dimensional vector
spaces over
F
, then there are bases (
e
1
, ··· , e
m
) for
U
and (
f
1
, ··· , f
n
) for
V
such that α is represented by the matrix
I
r
0
0 0
,
where r = r(α) and I
r
is the r × r identity matrix.
In particular, r(α) + n(α) = dim U.
Proof.
Let
e
k+1
, ··· , e
m
be a basis for the kernel of
α
. Then we can extend this
to a basis of the (e
1
, ··· , e
m
).
Let
f
i
=
α
(
e
i
) for 1
i k
. We now show that (
f
1
, ··· , f
k
) is a basis for
im α
(and thus
k
=
r
). We first show that it spans. Suppose
v im α
. Then we
have
v = α
m
X
i=1
λ
i
e
i
!
for some λ
i
F. By linearity, we can write this as
v =
m
X
i=1
λ
i
α(e
i
) =
k
X
i=1
λ
i
f
i
+ 0.
So v hf
1
, ··· , f
k
i.
To show linear dependence, suppose that
k
X
i=1
µ
i
f
i
= 0.
So we have
α
k
X
i=1
µ
i
e
i
!
= 0.
So
P
k
i=1
µ
i
e
i
ker α. Since (e
k+1
, ··· , e
m
) is a basis for ker α, we can write
k
X
i=1
µ
i
e
i
=
m
X
i=k+1
µ
i
e
i
for some
µ
i
(
i
=
k
+ 1
, ··· , m
). Since (
e
1
, ··· , e
m
) is a basis, we must have
µ
i
= 0 for all i. So they are linearly independent.
Now we extend (f
1
, ··· , f
r
) to a basis for V , and
α(e
i
) =
(
f
i
1 i k
0 k + 1 i m
.
Example. Let
W = {x R
5
: x
1
+ x
2
+ x
3
= 0 = x
3
x
4
x
5
}.
What is dim W ? Well, it clearly is 3, but how can we prove it?
We can consider the map α : R
5
R
2
given by
x
1
.
.
.
x
5
7→
x
1
+ x
2
+ x
5
x
3
x
4
x
5
.
Then
ker α
=
W
. So
dim W
= 5
r
(
α
). We know that
α
(1
,
0
,
0
,
0
,
0) = (1
,
0)
and α(0, 0, 1, 0, 0) = (0, 1). So r(α) = dim im α = 2. So dim W = 3.
More generally, the rank-nullity theorem gives that
m
linear equations in
n
have a space of solutions of dimension at least n m.
Example.
Suppose that
U
and
W
are subspaces of
V
, all of which are finite-
dimensional vector spaces of F. We let
α : U W V
(u, w) 7→ u + w,
where the is the external direct sum. Then im α = U + W and
ker α = {(u, u) : u U W }
=
dim(U W ).
Then we have
dim U + dim W = dim(U W ) = r(α) + n(α) = dim(U + W ) + dim(U W ).
This is a result we’ve previously obtained through fiddling with basis and horrible
stuff.
Corollary.
Suppose
α
:
U V
is a linear map between vector spaces over
F
both of dimension n < . Then the following are equivalent
(i) α is injective;
(ii) α is surjective;
(iii) α is an isomorphism.
Proof.
It is clear that, (iii) implies (i) and (ii), and (i) and (ii) together implies
(iii). So it suffices to show that (i) and (ii) are equivalent.
Note that
α
is injective iff
n
(
α
) = 0, and
α
is surjective iff
r
(
α
) =
dim V
=
n
.
By the rank-nullity theorem,
n
(
α
) +
r
(
α
) =
n
. So the result follows immediately.
Lemma.
Let
A M
n,n
(
F
) =
M
n
(
F
) be a square matrix. The following are
equivalent
(i) There exists B M
n
(F) such that BA = I
n
.
(ii) There exists C M
n
(F) such that AC = I
n
.
If these hold, then
B
=
C
. We call
A
invertible or non-singular, and write
A
1
= B = C.
Proof.
Let
α, β, γ, ι
:
F
n
F
n
be the linear maps represented by matrices
A, B, C, I
n
respectively with respect to the standard basis.
We note that (i) is equivalent to saying that there exists
β
such that
βα
=
ι
.
This is true iff
α
is injective, which is true iff
α
is an isomorphism, which is true
iff α has an inverse α
1
.
Similarly, (ii) is equivalent to saying that there exists
γ
such that
αγ
=
ι
.
This is true iff
α
is injective, which is true iff
α
is isomorphism, which is true iff
α has an inverse α
1
.
So these are the same things, and we have β = α
1
= γ.
2.4 Change of basis
Suppose we have a linear map
α
:
U V
. Given a basis
{e
i
}
for
U
, and a basis
{f
i
} for V , we can obtain a matrix A.
U V
F
m
F
n
α
A
(e
i
) (f
i
)
We now want to consider what happens when we have two different basis
{u
i
}
and
{e
i
}
of
U
. These will then give rise to two different maps from
F
m
to our
space
U
, and the two basis can be related by a change-of-basis map
P
. We can
put them in the following diagram:
U U
F
m
F
m
ι
U
P
(u
i
) (e
i
)
where
ι
U
is the identity map. If we perform a change of basis for both
U
and
V
,
we can stitch the diagrams together as
U U V V
F
m
F
m
F
n
F
n
ι
U α
ι
V
P
(u
i
)
B
A
(e
i
) (f
i
)
Q
(v
i
)
Then if we want a matrix representing the map
U V
with respect to bases
(u
i
) and (v
i
), we can write it as the composition
B = Q
1
AP.
We can write this as a theorem:
Theorem.
Suppose that (
e
1
, ··· , e
m
) and (
u
1
, ··· , u
m
) are basis for a finite-
dimensional vector space
U
over
F
, and (
f
1
, ··· , f
n
) and (
v
1
, ··· , v
n
) are basis
of a finite-dimensional vector space V over F.
Let
α
:
U V
be a linear map represented by a matrix
A
with respect to
(e
i
) and (f
i
) and by B with respect to (u
i
) and (v
i
). Then
B = Q
1
AP,
where P and Q are given by
u
i
=
m
X
k=1
P
ki
e
k
, v
i
=
n
X
k=1
Q
ki
f
k
.
Note that one can view
P
as the matrix representing the identity map
i
U
from
U
with basis (
u
i
) to
U
with basis (
e
i
), and similarly for
Q
. So both are
invertible.
Proof. On the one hand, we have
α(u
i
) =
n
X
j=1
B
ji
v
j
=
X
j
X
`
B
ji
Q
`j
f
`
=
X
`
[QB]
`i
f
`
.
On the other hand, we can write
α(u
i
) = α
m
X
k=1
P
ki
e
k
!
=
m
X
k=1
P
ki
X
`
A
`k
f
`
=
X
`
[AP ]
`i
f
`
.
Since the f
`
are linearly independent, we conclude that
QB = AP.
Since Q is invertible, we get B = Q
1
AP .
Definition
(Equivalent matrices)
.
We say
A, B Mat
n,m
(
F
) are equivalent if
there are invertible matrices
P Mat
m
(
F
),
Q Mat
n
(
F
) such that
B
=
Q
1
AP
.
Since
GL
K
(
F
) =
{A Mat
k
(
F
) :
A is invertible}
is a group, for each
k
1,
this is indeed an equivalence relation. The equivalence classes are orbits under
the action of GL
m
(F) × GL
n
(F), given by
GL
m
(F) × GL
n
(F) × Mat
n,m
(F) Mat(F)
(P, Q, A) 7→ QAP
1
.
Two matrices are equivalent if and only if they represent the same linear map
with respect to different basis.
Corollary.
If
A Mat
n,m
(
F
), then there exists invertible matrices
P
GL
m
(F), Q GL
n
(F) so that
Q
1
AP =
I
r
0
0 0
for some 0 r min(m, n).
This is just a rephrasing of the proposition we had last time. But this tells
us there are min(m, n) + 1 orbits of the action above parametrized by r.
Definition (Column and row rank). If A Mat
n,m
(F), then
The column rank of
A
, written
r
(
A
), is the dimension of the subspace of
F
n
spanned by the columns of A.
The row rank of
A
, written
r
(
A
), is the dimension of the subspace of
F
m
spanned by the rows of A. Alternatively, it is the column rank of A
T
.
There is no a priori reason why these should be equal to each other. However,
it turns out they are always equal.
Note that if
α
:
F
m
F
n
is the linear map represented by
A
(with respect to
the standard basis), then
r
(
A
) =
r
(
α
), i.e. the column rank is the rank. Moreover,
since the rank of a map is independent of the basis, equivalent matrices have
the same column rank.
Theorem.
If
A Mat
n,m
(
F
), then
r
(
A
) =
r
(
A
T
), i.e. the row rank is equivalent
to the column rank.
Proof. We know that there are some invertible P, Q such that
Q
1
AP =
I
r
0
0 0
,
where r = r(A). We can transpose this whole equation to obtain
(Q
1
AP )
T
= P
T
A
T
(Q
T
)
1
=
I
r
0
0 0
So r(A
T
) = r.
2.5 Elementary matrix operations
We are now going to re-prove our corollary that we can find
P, Q
such that
Q
1
AP
=
I
r
0
0 0
in a way that involves matrices only. This will give a
concrete way to find P and Q, but is less elegant.
To do so, we need to introduce elementary matrices.
Definition
(Elementary matrices)
.
We call the following matrices of
GL
n
(
F
)
elementary matrices:
S
n
ij
=
1
.
.
.
1
0 1
1
.
.
.
1
1 0
1
.
.
.
1
This is called a reflection, where the rows we changed are the ith and jth row.
E
n
ij
(λ) =
1
.
.
.
1 λ
.
.
.
1
.
.
.
1
This is called a shear, where λ appears at the i, jth entry.
T
n
i
(λ) =
1
.
.
.
1
λ
1
.
.
.
1
This is called a shear, where λ 6= 0 appears at the ith column.
Observe that if A is a m × n matrix, then
(i) AS
n
ij
is obtained from A by swapping the i and j columns.
(ii) AE
n
Ij
(λ) is obtained by adding λ× column i to column j.
(iii) AT
n
i
(λ) is obtained from A by rescaling the ith column by λ.
Multiplying on the left instead of the right would result in the same operations
performed on the rows instead of the columns.
Proposition.
If
A Mat
n,m
(
F
), then there exists invertible matrices
P
GL
m
(F), Q GL
n
(F) so that
Q
1
AP =
I
r
0
0 0
for some 0 r min(m, n).
We are going to start with
A
, and then apply these operations to get it into
this form.
Proof.
We claim that there are elementary matrices
E
m
1
, ··· , E
m
a
and
F
n
1
, ··· , F
n
b
(these E are not necessarily the shears, but any elementary matrix) such that
E
m
1
···E
m
a
AF
n
1
···F
n
b
=
I
r
0
0 0
This suffices since the
E
m
i
GL
M
(
F
) and
F
n
j
GL
n
(
F
). Moreover, to prove the
claim, it suffices to find a sequence of elementary row and column operations
reducing A to this form.
If
A
= 0, then done. If not, there is some
i, j
such that
A
ij
6
= 0. By swapping
row 1 and row
i
; and then column 1 and column
j
, we can assume
A
11
6
= 0. By
rescaling row 1 by
1
A
11
, we can further assume A
11
= 1.
Now we can add
A
1j
times column 1 to column
j
for each
j 6
= 1, and then
add A
i1
times row 1 to row i 6= 1. Then we now have
A =
1 0 ··· 0
0
.
.
. B
0
Now
B
is smaller than
A
. So by induction on the size of
A
, we can reduce
B
to
a matrix of the required form, so done.
It is an exercise to show that the row and column operations do not change
the row rank or column rank, and deduce that they are equal.
3 Duality
Duality is a principle we will find throughout mathematics. For example, in IB
Optimisation, we considered the dual problems of linear programs. Here we will
look for the dual of vector spaces. In general, we try to look at our question in a
“mirror” and hope that the mirror problem is easier to solve than the original
mirror.
At first, the definition of the dual might see a bit arbitrary and weird. We
will try to motivate it using what we will call annihilators, but they are much
more useful than just for these. Despite their usefulness, though, they can be
confusing to work with at times, since the dual space of a vector space
V
will be
constructed by considering linear maps on
V
, and when we work with maps on
dual spaces, things explode.
3.1 Dual space
To specify a subspace of
F
n
, we can write down linear equations that its elements
satisfy. For example, if we have the subspace
U
=
h
1
2
1
i F
3
, we can specify
this by saying
x
1
x
2
x
3
U if and only if
x
1
x
3
= 0
2x
1
x
2
= 0.
However, characterizing a space in terms of equations involves picking some
particular equations out of the many possibilities. In general, we do not like
making arbitrary choices. Hence the solution is to consider all possible such
equations. We will show that these form a subspace in some space.
We can interpret these equations in terms of linear maps
F
n
F
. For
example
x
1
x
3
= 0 if and only if
x
1
x
2
x
3
ker θ
, where
θ
:
F
3
F
is defined
by
x
1
x
2
x
3
7→ x
1
x
3
.
This works well with the vector space operations. If
θ
1
, θ
2
:
F
n
F
vanish on
some subspace of
F
n
, and
λ, µ F
, then
λθ
1
+
µθ
2
also vanishes on the subspace.
So the set of all maps F
n
F that vanishes on U forms a vector space.
To formalize this notion, we introduce dual spaces.
Definition
(Dual space)
.
Let
V
be a vector space over
F
. The dual of
V
is
defined as
V
= L(V, F) = {θ : V F : θ linear}.
Elements of V
are called linear functionals or linear forms.
By convention, we use Roman letters for elements in
V
, and Greek letters
for elements in V
.
Example.
If V = R
3
and θ : V R that sends
x
1
x
2
x
3
7→ x
1
x
3
, then θ V
.
Let
V
=
F
X
. Then for any fixed
x
,
θ
:
V F
defined by
f 7→ f
(
x
) is in
V
.
Let V = C([0, 1], R). Then f 7→
R
1
0
f(t) dt V
.
The trace tr : M
n
(F) F defined by A 7→
P
n
i=1
A
ii
is in M
n
(F)
.
It turns out it is rather easy to specify how the dual space looks like, at least
in the case where V is finite dimensional.
Lemma.
If
V
is a finite-dimensional vector space over
f
with basis (
e
1
, ··· , e
n
),
then there is a basis (
ε
1
, ··· , ε
n
) for
V
(called the dual basis to (
e
1
, ··· , e
n
))
such that
ε
i
(e
j
) = δ
ij
.
Proof.
Since linear maps are characterized by their values on a basis, there exists
unique choices for ε
1
, ··· , ε
n
V
. Now we show that (ε
1
, ··· , ε
n
) is a basis.
Suppose
θ V
. We show that we can write it uniquely as a combination of
ε
1
, ··· , ε
n
. We have
θ
=
P
n
i=1
λ
i
ε
i
if and only if
θ
(
e
j
) =
P
n
i=1
λ
i
ε
i
(
e
j
) (for all
j) if and only if λ
j
= θ(e
j
). So we have uniqueness and existence.
Corollary. If V is finite dimensional, then dim V = dim V
.
When
V
is not finite dimensional, this need not be true. However, we
know that the dimension of
V
is at least as big as that of
V
, since the above
gives a set of
dim V
many independent vectors in
V
. In fact for any infinite
dimensional vector space,
dim V
is strictly larger than
dim V
, if we manage to
define dimensions for infinite-dimensional vector spaces.
It helps to come up with a more concrete example of how dual spaces look
like. Consider the vector space
F
n
, where we treat each element as a column
vector (with respect to the standard basis). Then we can regard elements of
V
as just row vectors (
a
1
, ··· , a
n
) =
P
n
j=1
a
j
ε
j
with respect to the dual basis. We
have
X
a
j
ε
j
X
x
i
e
i
!
=
X
i,j
a
j
x
i
δ
ij
=
n
X
i=1
a
i
x
i
=
a
1
··· a
n
x
1
.
.
.
x
n
.
This is exactly what we want.
Now what happens when we change basis? How will the dual basis change?
Proposition.
Let
V
be a finite-dimensional vector space over
F
with bases
(e
1
, ··· , e
n
) and (f
1
, ··· , f
n
), and that P is the change of basis matrix so that
f
i
=
n
X
k=1
P
ki
e
k
.
Let (ε
1
, ··· , ε
n
) and (η
1
, ··· , η
n
) be the corresponding dual bases so that
ε
i
(e
j
) = δ
ij
= η
i
(f
j
).
Then the change of basis matrix from (
ε
1
, ··· , ε
n
) to (
η
1
, ··· , η
n
) is (
P
1
)
T
, i.e.
ε
i
=
n
X
`=1
P
T
`i
η
`
.
Proof. For convenience, write Q = P
1
so that
e
j
=
n
X
k=1
Q
kj
f
k
.
So we can compute
n
X
`=1
P
i`
η
`
!
(e
j
) =
n
X
`=1
P
i`
η
`
!
n
X
k=1
Q
kj
f
k
!
=
X
k,`
P
i`
δ
`k
Q
kj
=
X
k,`
P
i`
Q
`j
= [P Q]
ij
= δ
ij
.
So ε
i
=
P
n
`=1
P
T
`i
η
`
.
Now we’ll return to our original motivation, and think how we can define
subspaces of V
in terms of subspaces of V , and vice versa.
Definition (Annihilator). Let U V . Then the annihilator of U is
U
0
= {θ V
: θ(u) = 0, u U}.
If W V
, then the annihilator of W is
W
0
= {v V : θ(v) = 0, θ W }.
One might object that
W
0
should be a subset of
V
∗∗
and not
V
. We will
later show that there is a canonical isomorphism between
V
∗∗
and
V
, and this
will all make sense.
Example.
Consider
R
3
with standard basis
e
1
, e
2
, e
3
; (
R
3
)
with dual basis
(
ε
1
, ε
2
, ε
3
). If
U
=
he
1
+ 2
e
2
+
e
3
i
and
W
=
hε
1
ε
3
,
2
ε
1
ε
2
i
, then
U
0
=
W
and W
0
= U.
We see that the dimension of
U
and
U
0
add up to three, which is the
dimension of R
3
. This is typical.
Proposition. Let V be a vector space over F and U a subspace. Then
dim U + dim U
0
= dim V.
We are going to prove this in many ways.
Proof.
Let (
e
1
, ··· , e
k
) be a basis for
U
and extend to (
e
1
, ··· , e
n
) a basis for
V . Consider the dual basis for V
, say (ε
1
, ··· , ε
n
). Then we will show that
U
0
= hε
k+1
, ··· , ε
n
i.
So
dim U
0
=
n k
as required. This is easy to prove if
j > k
, then
ε
j
(
e
i
) = 0
for all
i k
. So
ε
k+1
, ··· , ε
n
U
0
. On the other hand, suppose
θ U
0
. Then
we can write
θ =
n
X
j=1
λ
j
ε
j
.
But then 0 = θ(e
i
) = λ
i
for i k. So done.
Proof.
Consider the restriction map
V
U
, given by
θ 7→ θ|
U
. This is
obviously linear. Since every linear map
U F
can be extended to
V F
, this
is a surjection. Moreover, the kernel is U
0
. So by rank-nullity theorem,
dim V
= dim U
0
+ dim U
.
Since dim V
= dim V and dim U
= dim U, we’re done.
Proof.
We can show that
U
0
'
(
V/U
)
, and then deduce the result. Details are
left as an exercise.
3.2 Dual maps
Since linear algebra is the study of vector spaces and linear maps between them,
after dualizing vector spaces, we should be able to dualize linear maps as well.
If we have a map
α
:
V W
, then after dualizing, the map will go the other
direction, i.e.
α
:
W
V
. This is a characteristic common to most dualization
processes in mathematics.
Definition
(Dual map)
.
Let
V, W
be vector spaces over
F
and
α
:
V W
L
(
V, W
). The dual map to
α
, written
α
:
W
V
is given by
θ 7→ θ α
.
Since the composite of linear maps is linear,
α
(
θ
)
V
. So this is a genuine
map.
Proposition.
Let
α L
(
V, W
) be a linear map. Then
α
L
(
W
, V
) is a
linear map.
This is not the same as what we remarked at the end of the definition of
the dual map. What we remarked was that given any
θ
,
α
(
θ
) is a linear map.
What we want to show here is that α
itself as a map W
V
is linear.
Proof. Let λ, µ F and θ
1
, θ
2
W
. We want to show
α
(λθ
1
+ µθ
2
) = λα
(θ
1
) + µα
(θ
2
).
To show this, we show that for every
v V
, the left and right give the same
result. We have
α
(λθ
1
+ µθ
2
)(v) = (λθ
1
+ µθ
2
)(αv)
= λθ
1
(α(v)) + µθ
2
(α(v))
= (λα
(θ
1
) + µα
(θ
2
))(v).
So α
L(W
, V
).
What happens to the matrices when we take the dual map? The answer is
that we get the transpose.
Proposition.
Let
V, W
be finite-dimensional vector spaces over
F
and
α
:
V
W
be a linear map. Let (
e
1
, ··· , e
n
) be a basis for
V
and (
f
1
, ··· , f
m
) be a basis
for W ; (ε
1
, ··· , ε
n
) and (η
1
, ··· , η
m
) the corresponding dual bases.
Suppose
α
is represented by
A
with respect to (
e
i
) and (
f
i
) for
V
and
W
.
Then α
is represented by A
T
with respect to the corresponding dual bases.
Proof. We are given that
α(e
i
) =
m
X
k=1
A
ki
f
k
.
We must compute α
(η
i
). To do so, we evaluate it at e
j
. We have
α
(η
i
)(e
j
) = η
i
(α(e
j
)) = η
i
m
X
k=1
A
kj
f
k
!
=
m
X
k=1
A
kj
δ
ik
= A
ij
.
We can also write this as
α
(η
i
)(e
j
) =
n
X
k=1
A
ik
ε
k
(e
j
).
Since this is true for all j, we have
α
(η
i
)
n
X
k=1
A
ik
ε
k
=
n
X
k=1
A
T
ki
ε
k
.
So done.
Note that if α : U V and β : V W , θ W
, then
(βα)
(θ) = θβα = α
(θβ) = α
(β
(θ)).
So we have (
βα
)
=
α
β
. This is obviously true for the finite-dimensional case,
since that’s how transposes of matrices work.
Similarly, if α, β : U V , then (λα + µβ)
= λα
+ µβ
.
What happens when we change basis? If
B
=
Q
1
AP
for some invertible
P
and Q, then
B
T
= (Q
1
AP )
T
= P
T
A
T
(Q
1
)
T
= ((P
1
)
T
)
1
A
T
(Q
1
)
T
.
So in the dual space, we conjugate by the dual of the change-of-basis matrices.
As we said, we can use dualization to translate problems about a vector space
to its dual. The following lemma gives us some good tools to do so:
Lemma.
Let
α L
(
V, W
) with
V, W
finite dimensional vector spaces over
F
.
Then
(i) ker α
= (im α)
0
.
(ii) r
(
α
) =
r
(
α
) (which is another proof that row rank is equal to column
rank).
(iii) im α
= (ker α)
0
.
At first sight, (i) and (iii) look quite similar. However, (i) is almost trivial to
prove, but (iii) is rather hard.
Proof.
(i) If θ W
, then
θ ker α
α
(θ) = 0
(v V ) θα(v) = 0
(w im α) θ(w) = 0
θ (im α)
0
.
(ii) As im α W , we’ve seen that
dim im α + dim(im α)
0
= dim W.
Using (i), we see
n(α
) = dim(im α)
0
.
So
r(α) + n(α
) = dim W = dim W
.
By the rank-nullity theorem, we have r(α) = r(α
).
(iii)
The proof in (i) doesn’t quite work here. We can only show that one
includes the other. To draw the conclusion, we will show that the two
spaces have the dimensions, and hence must be equal.
Let θ im α
. Then θ = φα for some φ W
. If v ker α, then
θ(v) = φ(α(v)) = φ(0) = 0.
So im α
(ker α)
0
.
But we know
dim(ker α)
0
+ dim ker α = dim V,
So we have
dim(ker α)
0
= dim V n(α) = r(α) = r(α
) = dim im α
.
Hence we must have im α
= (ker α)
0
.
Not only do we want to get from
V
to
V
, we want to get back from
V
to
V
.
We can take the dual of
V
to get a
V
∗∗
. We already know that
V
∗∗
is isomorphic
to
V
, since
V
is isomorphic to
V
already. However, the isomorphism between
V
and
V
are not “natural”. To define such an isomorphism, we needed to pick
a basis for
V
and consider a dual basis. If we picked a different basis, we would
get a different isomorphism. There is no natural, canonical, uniquely-defined
isomorphism between V and V
.
However, this is not the case when we want to construct an isomorphism
V V
∗∗
. The construction of this isomorphism is obvious once we think hard
what
V
∗∗
actually means. Unwrapping the definition, we know
V
∗∗
=
L
(
V
, F
).
Our isomorphism has to produce something in
V
∗∗
given any
v V
. This is
equivalent to saying given any
v V
and a function
θ V
, produce something
in F.
This is easy, by definition
θ V
is just a linear map
V F
. So given
v
and θ, we just return θ(v). We now just have to show that this is linear and is
bijective.
Lemma.
Let
V
be a vector space over
F
. Then there is a linear map
ev
:
V
(V
)
given by
ev(v)(θ) = θ(v).
We call this the evaluation map.
We call this a “canonical” map since this does not require picking a particular
basis of the vector spaces. It is in some sense a “natural” map.
Proof.
We first show that
ev
(
v
)
V
∗∗
for all
v V
, i.e.
ev
(
v
) is linear for any
v. For any λ, µ F, θ
1
, θ
2
V
, then for v V , we have
ev(v)(λθ
1
+ µθ
2
) = (λθ
1
+ µθ
2
)(v)
= λθ
1
(v) + µθ
2
(v)
= λ ev(v)(θ
1
) + µ ev(v)(θ
2
).
So done. Now we show that
ev
itself is linear. Let
λ, µ F
,
v
1
, v
2
V
. We
want to show
ev(λv
1
+ µv
2
) = λ ev(v
1
) + µ ev(v
2
).
To show these are equal, pick θ V
. Then
ev(λv
1
+ µv
2
)(θ) = θ(λv
1
+ µv
2
)
= λθ(v
1
) + µθ(v
2
)
= λ ev(v
1
)(θ) + µ ev(v
2
)(θ)
= (λ ev(v
1
) + µ ev(v
2
))(θ).
So done.
In the special case where V is finite-dimensional, this is an isomorphism.
Lemma. If V is finite-dimensional, then ev : V V
∗∗
is an isomorphism.
This is very false for infinite dimensional spaces. In fact, this is true only
for finite-dimensional vector spaces (assuming the axiom of choice), and some
(weird) people use this as the definition of finite-dimensional vector spaces.
Proof.
We first show it is injective. Suppose
ev
(
v
) =
0
for some
v V
. Then
θ
(
v
) =
ev
(
v
)(
θ
) = 0 for all
θ V
. So
dimhvi
0
=
dim V
=
dim V
. So
dimhvi
= 0. So
v
= 0. So
ev
is injective. Since
V
and
V
∗∗
have the same
dimension, this is also surjective. So done.
From now on, we will just pretend that
V
and
V
∗∗
are the same thing, at
least when V is finite dimensional.
Note that this lemma does not just say that
V
is isomorphic to
V
∗∗
(we
already know that since they have the same dimension). This says there is a
completely canonical way to choose the isomorphism.
In general, if
V
is infinite dimensional, then
ev
is injective, but not surjective.
So we can think of V as a subspace of V
∗∗
in a canonical way.
Lemma.
Let
V, W
be finite-dimensional vector spaces over
F
after identifying
(V and V
∗∗
) and (W and W
∗∗
) by the evaluation map. Then we have
(i) If U V , then U
00
= U.
(ii) If α L(V, W ), then α
∗∗
= α.
Proof.
(i)
Let
u U
. Then
u
(
θ
) =
θ
(
u
) = 0 for all
θ U
0
. So
u
annihilates
everything in U
0
. So u U
00
. So U U
00
. We also know that
dim U = dim V dim U
0
= dim V (dim V dim U
00
) = dim U
00
.
So we must have U = U
00
.
(ii)
The proof of this is basically the transpose of the transpose is the
original matrix. The only work we have to do is to show that the dual of
the dual basis is the original basis.
Let (
e
1
, ··· , e
n
) be a basis for
V
and (
f
1
, ··· , f
m
) be a basis for
W
, and let
(
ε
1
, ··· , ε
n
) and (
η
1
, ··· , η
n
) be the corresponding dual basis. We know
that
e
i
(ε
j
) = δ
ij
= ε
j
(e
i
), f
i
(η
j
) = δ
ij
= η
j
(f
i
).
So (e
1
, ··· , e
n
) is dual to (ε
1
, ··· , ε
n
), and similarly for f and η.
If
α
is represented by
A
, then
α
is represented by
A
T
. So
α
∗∗
is represented
by (A
T
)
T
= A. So done.
Proposition.
Let
V
be a finite-dimensional vector space
F
and
U
1
,
U
2
are
subspaces of V . Then we have
(i) (U
1
+ U
2
)
0
= U
0
1
U
0
2
(ii) (U
1
U
2
)
0
= U
0
1
+ U
0
2
Proof.
(i) Suppose θ V
. Then
θ (U
1
+ U
2
)
0
θ(u
1
+ u
2
) = 0 for all u
i
U
i
θ(u) = 0 for all u U
1
U
2
θ U
0
1
U
0
2
.
(ii) We have
(U
1
U
2
)
0
= ((U
0
1
)
0
(U
0
2
)
0
)
0
= (U
0
1
+ U
0
2
)
00
= U
0
1
+ U
0
2
.
So done.
4 Bilinear forms I
So far, we have been looking at linear things only. This can get quite boring.
For a change, we look at bilinear maps instead. In this chapter, we will look at
bilinear forms in general. It turns out there isn’t much we can say about them,
and hence this chapter is rather short. Later, in Chapter 7, we will study some
special kinds of bilinear forms which are more interesting.
Definition
(Bilinear form)
.
Let
V, W
be vector spaces over
F
. Then a function
φ
:
V × W F
is a bilinear form if it is linear in each variable, i.e. for each
v V , φ(v, ·) : W F is linear; for each w W , φ( ·, w) : V F is linear.
Example. The map defined by
V × V
F
(v, θ) 7→ θ(v) = ev(v)(θ)
is a bilinear form.
Example.
Let
V
=
W
=
F
n
. Then the function (
v, w
) =
P
n
i=1
v
i
w
i
is bilinear.
Example. If V = W = C([0, 1], R), then
(f, g) 7→
Z
a
0
fg dt
is a bilinear form.
Example. Let A Mat
m,n
(F). Then
φ : F
m
× F
n
F
(v, w) 7→ v
T
Aw
is bilinear. Note that the (real) dot product is the special case of this, where
n = m and A = I.
In fact, this is the most general form of bilinear forms on finite-dimensional
vector spaces.
Definition
(Matrix representing bilinear form)
.
Let (
e
1
, ··· , e
n
) be a basis for
V
and (
f
1
, ··· , f
m
) be a basis for
W
, and
ψ
:
V × W F
. Then the matrix
A
representing ψ with respect to the basis is defined to be
A
ij
= ψ(e
i
, f
j
).
Note that if v =
P
λ
i
e
i
and w =
P
µ
j
f
j
, then by linearity, we get
ψ(v, w) = ψ
X
λ
i
e
i
, w
=
X
i
λ
i
ψ(e
i
, w)
=
X
i
λ
i
ψ
e
i
,
X
µ
j
f
j
=
X
i,j
λ
i
µ
j
ψ(e
i
, f
j
)
= λ
T
Aµ.
So ψ is determined by A.
We have identified linear maps with matrices, and we have identified bilinear
maps with matrices. However, you shouldn’t think linear maps are bilinear maps.
They are, obviously, two different things. In fact, the matrices representing
matrices and bilinear forms transform differently when we change basis.
Proposition.
Suppose (
e
1
, ··· , e
n
) and (
v
1
, ··· , v
n
) are basis for
V
such that
v
i
=
X
P
ki
e
k
for all i = 1, ··· , n;
and (f
1
, ··· , f
m
) and (w
1
, ··· , w
m
) are bases for W such that
w
i
=
X
Q
`j
f
`
for all j = 1, ··· , m.
Let
ψ
:
V × W F
be a bilinear form represented by
A
with respect to
(
e
1
, ··· , e
n
) and (
f
1
, ··· , f
m
), and by
B
with respect to the bases (
v
1
, ··· , v
n
)
and (w
1
, ··· , w
m
). Then
B = P
T
AQ.
The difference with the transformation laws of matrices is this time we are
taking transposes, not inverses.
Proof. We have
B
ij
= φ(v
i
, w
j
)
= φ
X
P
ki
e
k
,
X
Q
`j
f
`
=
X
P
ki
Q
`j
φ(e
k
, f
`
)
=
X
k,`
P
T
ik
A
k`
Q
`j
= (P
T
AQ)
ij
.
Note that while the transformation laws for bilinear forms and linear maps
are different, we still get that two matrices are representing the same bilinear
form with respect to different bases if and only if they are equivalent, since if
B = P
1
AQ, then B = ((P
1
)
T
)
T
AQ.
If we are given a bilinear form
ψ
:
V ×W F
, we immediately get two linear
maps:
ψ
L
: V W
, ψ
R
: W V
,
defined by ψ
L
(v) = ψ(v, ·) and ψ
R
(w) = ψ( ·, w).
For example, if
ψ
:
V × V
F
, is defined by (
v, θ
)
7→ θ
(
v
), then
ψ
L
:
V
V
∗∗
is the evaluation map. On the other hand,
ψ
R
:
V
V
is the identity
map.
Lemma.
Let (
ε
1
, ··· , ε
n
) be a basis for
V
dual to (
e
1
, ··· , e
n
) of
V
; (
η
1
, ··· , η
n
)
be a basis for W
dual to (f
1
, ··· , f
n
) of W .
If
A
represents
ψ
with respect to (
e
1
, ··· , e
n
) and (
f
1
, ··· , f
m
), then
A
also
represents
ψ
R
with respect to (
f
1
, ··· , f
m
) and (
ε
1
, ··· , ε
n
); and
A
T
represents
ψ
L
with respect to (e
1
, ··· , e
n
) and (η
1
, ··· , η
m
).
Proof. We just have to compute
ψ
L
(e
i
)(f
j
) = A
ij
=
X
A
i`
η
`
(f
j
).
So we get
ψ
L
(e
i
) =
X
A
T
`i
η
`
.
So A
T
represents ψ
L
.
We also have
ψ
R
(f
j
)(e
i
) = A
ij
.
So
ψ
R
(f
j
) =
X
A
kj
ε
k
.
Definition
(Left and right kernel)
.
The kernel of
ψ
L
is left kernel of
ψ
, while
the kernel of ψ
R
is the right kernel of ψ.
Then by definition, v is in the left kernel if ψ(v, w) = 0 for all w W .
More generally, if T V , then we write
T
= {w W : ψ(t, w) = 0 for all t T }.
Similarly, if U W , then we write
U = {v V : ψ(v, u) = 0 for all u U}.
In particular, V
= ker ψ
R
and
W = ker ψ
L
.
If we have a non-trivial left (or right) kernel, then in some sense, some
elements in V (or W ) are “useless”, and we don’t like these.
Definition
(Non-degenerate bilinear form)
. ψ
is non-degenerate if the left and
right kernels are both trivial. We say ψ is degenerate otherwise.
Definition
(Rank of bilinear form)
.
If
ψ
:
V W
is a bilinear form
F
on a
finite-dimensional vector space
V
, then the rank of
V
is the rank of any matrix
representing
φ
. This is well-defined since
r
(
P
T
AQ
) =
r
(
A
) if
P
and
Q
are
invertible.
Alternatively, it is the rank of ψ
L
(or ψ
R
).
Lemma.
Let
V
and
W
be finite-dimensional vector spaces over
F
with bases
(e
1
, ··· , e
n
) and (f
1
, ··· , f
m
) be their basis respectively.
Let
ψ
:
V ×W F
be a bilinear form represented by
A
with respect to these
bases. Then
φ
is non-degenerate if and only if
A
is (square and) invertible. In
particular, V and W have the same dimension.
We can understand this as saying if there are too many things in
V
(or
W
),
then some of them are bound to be useless.
Proof.
Since
ψ
R
and
ψ
L
are represented by
A
and
A
T
(in some order), they both
have trivial kernel if and only if
n
(
A
) =
n
(
A
T
) = 0. So we need
r
(
A
) =
dim V
and
r
(
A
T
) =
dim W
. So we need
dim V
=
dim W
and
A
have full rank, i.e. the
corresponding linear map is bijective. So done.
Example. The map
F
2
× F
2
F
a
c
,
b
d
7→ ad bc
is a bilinear form. This, obviously, corresponds to the determinant of a 2-by-2
matrix. We have ψ(v, w) = ψ(w, v) for all v, w F
2
.
5 Determinants of matrices
We probably all know what the determinant is. Here we are going to give a
slightly more abstract definition, and spend quite a lot of time trying motivate
this definition.
Recall that
S
n
is the group of permutations of
{
1
, ··· , n}
, and there is a
unique group homomorphism
ε
:
S
n
1
}
such that
ε
(
σ
) = 1 if
σ
can be
written as a product of an even number of transpositions;
ε
(
σ
) =
1 if
σ
can be
written as an odd number of transpositions. It is proved in IA Groups that this
is well-defined.
Definition (Determinant). Let A Mat
n,n
(F). Its determinant is
det A =
X
σS
n
ε(σ)
n
Y
i=1
A
(i)
.
This is a big scary definition. Hence, we will spend the first half of the
chapter trying to understand what this really means, and how it behaves. We
will eventually prove a formula that is useful for computing the determinant,
which is probably how you were first exposed to the determinant.
Example. If n = 2, then S
2
= {id, (1 2)}. So
det A = A
11
A
22
A
12
A
21
.
When n = 3, then S
3
has 6 elements, and
det A = A
11
A
22
A
33
+ A
12
A
23
A
31
+ A
13
A
21
A
32
A
11
A
23
A
32
A
22
A
31
A
13
A
33
A
12
A
21
.
We will first prove a few easy and useful lemmas about the determinant.
Lemma. det A = det A
T
.
Proof.
det A
T
=
X
σS
n
ε(σ)
n
Y
i=1
A
σ(i)i
=
X
σS
n
ε(σ)
n
Y
j=1
A
jσ
1
(j)
=
X
τ S
n
ε(τ
1
)
n
Y
j=1
A
jτ (j)
Since ε(τ ) = ε(τ
1
), we get
=
X
τ S
n
ε(τ)
n
Y
j=1
A
jτ (j)
= det A.
Lemma. If A is an upper triangular matrix, i.e.
A =
a
11
a
12
··· a
1n
0 a
22
··· a
2n
.
.
.
.
.
.
.
.
.
.
.
.
0 0 ··· a
nn
Then
det A =
n
Y
i=1
a
ii
.
Proof. We have
det A =
X
σS
n
ε(σ)
n
Y
i=1
A
( i)
But A
( i)
= 0 whenever i > σ(i). So
n
Y
i=1
A
( i)
= 0
if there is some i {1, ··· , n} such that i > σ(i).
However, the only permutation in which
i σ
(
i
) for all
i
is the identity. So
the only thing that contributes in the sum is σ = id. So
det A =
n
Y
i=1
A
ii
.
To motivate this definition, we need a notion of volume. How can we define
volume on a vector space? It should be clear that the “volume” cannot be
uniquely determined, since it depends on what units we are using. For example,
saying the volume is “1” is meaningless unless we provide the units, e.g.
cm
3
.
So we have an axiomatic definition for what it means for something to denote a
“volume”.
Definition
(Volume form)
.
A volume form on
F
n
is a function
d
:
F
n
×···×F
n
F that is
(i) Multilinear, i.e. for all i and all v
1
, ··· , v
i1
, v
i+1
, ··· , v
n
F
n
, we have
d(v
1
, ··· , v
i1
, ·, v
i+1
, ··· , v
n
) (F
n
)
.
(ii) Alternating, i.e. if v
i
= v
j
for some i 6= j, then
d(v
1
, ··· , v
n
) = 0.
We should think of
d
(
v
1
, ··· , v
n
) as the
n
-dimensional volume of the paral-
lelopiped spanned by v
1
, ··· , v
n
.
We can view
A Mat
n
(
F
) as
n
-many vectors in
F
n
by considering its columns
A = (A
(1)
A
(2)
··· A
(n)
), with A
(i)
F
n
. Then we have
Lemma. det A is a volume form.
Proof. To see that det is multilinear, it is sufficient to show that each
n
Y
i=1
A
(i)
is multilinear for all
σ S
n
, since linear combinations of multilinear forms are
multilinear. But each such product is contains precisely one entry from each
column, and so is multilinear.
To show it is alternating, suppose now there are some
k, `
distinct such that
A
(k)
=
A
(`)
. We let
τ
be the transposition (
k `
). By Lagrange’s theorem, we
can write
S
n
= A
n
q τ A
n
,
where A
n
= ker ε and q is the disjoint union. We also know that
X
σA
n
n
Y
i=1
A
( i)
=
X
σA
n
n
Y
i=1
A
i,τ σ(i)
,
since if
σ
(
i
) is not
k
or
l
, then
τ
does nothing; if
σ
(
i
) is
k
or
l
, then
τ
just swaps
them around, but A
(k)
= A
(l)
. So we get
X
σA
n
n
Y
i=1
A
(i)
=
X
σ
0
τ A
n
n
Y
i=1
A
0
(i)
,
But we know that
det A = LHS RHS = 0.
So done.
We have shown that determinants are volume forms, but is this the only
volume form? Well obviously not, since 2
det A
is also a valid volume form.
However, in some sense, all volume forms are “derived” from the determinant.
Before we show that, we need the following
Lemma.
Let
d
be a volume form on
F
n
. Then swapping two entries changes
the sign, i.e.
d(v
1
, ··· , v
i
, ··· , v
j
, ··· , v
n
) = d(v
1
, ··· , v
j
, ··· , v
i
, ··· , v
n
).
Proof. By linearity, we have
0 = d(v
1
, ··· , v
i
+ v
j
, ··· , v
i
+ v
j
, ··· , v
n
)
= d(v
1
, ··· , v
i
, ··· , v
i
, ··· , v
n
)
+ d(v
1
, ··· , v
i
, ··· , v
j
, ··· , v
n
)
+ d(v
1
, ··· , v
j
, ··· , v
i
, ··· , v
n
)
+ d(v
1
, ··· , v
j
, ··· , v
j
, ··· , v
n
)
= d(v
1
, ··· , v
i
, ··· , v
j
, ··· , v
n
)
+ d(v
1
, ··· , v
j
, ··· , v
i
, ··· , v
n
).
So done.
Corollary. If σ S
n
, then
d(v
σ(1)
, ··· , v
σ(n)
) = ε(σ)d(v
1
, ··· , v
n
)
for any v
i
F
n
.
Theorem.
Let
d
be any volume form on
F
n
, and let
A
= (
A
(1)
··· A
(n)
)
Mat
n
(F). Then
d(A
(1)
, ··· , A
(n)
) = (det A)d(e
1
, ··· , e
n
),
where {e
1
, ··· , e
n
} is the standard basis.
Proof. We can compute
d(A
(1)
, ··· , A
(n)
) = d
n
X
i=1
A
i1
e
i
, A
(2)
, ··· , A
(n)
!
=
n
X
i=1
A
i1
d(e
i
, A
(2)
, ··· , A
(n)
)
=
n
X
i,j=1
A
i1
A
j2
d(e
i
, e
j
, A
(3)
, ··· , A
(n)
)
=
X
i
1
,··· ,i
n
d(e
i
1
, ··· , e
i
n
)
n
Y
j=1
A
i
j
j
.
We know that lots of these are zero, since if
i
k
=
i
j
for some
k, j
, then the term
is zero. So we are just summing over distinct tuples, i.e. when there is some
σ
such that i
j
= σ(j). So we get
d(A
(1)
, ··· , A
(n)
) =
X
σS
n
d(e
σ(1)
, ··· , e
σ(n)
)
n
Y
j=1
A
σ(j)j
.
However, by our corollary up there, this is just
d(A
(1)
, ··· , A
(n)
) =
X
σS
n
ε(σ)d(e
1
, ··· , e
n
)
n
Y
j=1
A
σ(j)j
= (det A)d(e
1
, ··· , e
n
).
So done.
We can rewrite the formula as
d(Ae
1
, ··· , Ae
n
) = (det A)d(e
1
, ··· , e
n
).
It is not hard to see that the same proof gives for any v
1
, ··· , v
n
, we have
d(Av
1
, ··· , Av
n
) = (det A)d(v
1
, ··· , v
n
).
So we know that
det A
is the volume rescaling factor of an arbitrary parallelopiped,
and this is true for any volume form d.
Theorem. Let A, B Mat
n
(F). Then det(AB) = det(A) det(B).
Proof.
Let
d
be a non-zero volume form on
F
n
(e.g. the “determinant”). Then
we can compute
d(ABe
1
, ··· , ABe
n
) = (det AB)d(e
1
, ··· , e
n
),
but we also have
d(ABe
1
, ··· , ABe
n
) = (det A)d(Be
1
, ··· , Be
n
) = (det A)(det B)d(e
1
, ··· , e
n
).
Since d is non-zero, we must have det AB = det A det B.
Corollary.
If
A Mat
n
(
F
) is invertible, then
det A 6
= 0. In fact, when
A
is
invertible, then det(A
1
) = (det A)
1
.
Proof. We have
1 = det I = det(AA
1
) = det A det A
1
.
So done.
Definition
(Singular matrices)
.
A matrix
A
is singular if
det A
= 0. Otherwise,
it is non-singular.
We have just shown that if
det A
= 0, then
A
is not invertible. Is the converse
true? If
det A 6
= 0, then can we conclude that
A
is invertible? The answer is
yes. We are now going to prove it in an abstract and clean way. We will later
prove this fact again by constructing an explicit formula for the inverse, which
involves dividing by the determinant. So if the determinant is non-zero, then we
know an inverse exists.
Theorem. Let A Mat
n
(F). Then the following are equivalent:
(i) A is invertible.
(ii) det A 6= 0.
(iii) r(A) = n.
Proof.
We have proved that (i)
(ii) above, and the rank-nullity theorem implies
(iii)
(i). We will prove (ii)
(iii). In fact we will show the contrapositive.
Suppose
r
(
A
)
< n
. By rank-nullity theorem,
n
(
A
)
>
0. So there is some
x =
λ
1
.
.
.
λ
n
such that Ax = 0. Suppose λ
k
6= 0. We define B as follows:
B =
1 λ
1
.
.
.
.
.
.
1 λ
k1
λ
k
λ
k+1
1
.
.
.
.
.
.
λ
n
1
So
AB
has the
k
th column identically zero. So
det
(
AB
) = 0. So it is sufficient
to prove that det(B) 6= 0. But det B = λ
k
6= 0. So done.
We are now going to come up with an alternative formula for the determinant
(which is probably the one you are familiar with). To do so, we introduce the
following notation:
Notation.
Write
ˆ
A
ij
for the matrix obtained from
A
by deleting the
i
th row
and jth column.
Lemma. Let A Mat
n
(F). Then
(i) We can expand det A along the jth column by
det A =
n
X
i=1
(1)
i+j
A
ij
det
ˆ
A
ij
.
(ii) We can expand det A along the ith row by
det A =
n
X
j=1
(1)
i+j
A
ij
det
ˆ
A
ij
.
We could prove this directly from the definition, but that is messy and scary,
so let’s use volume forms instead.
Proof.
Since
det A
=
det A
T
, (i) and (ii) are equivalent. So it suffices to prove
just one of them. We have
det A = d(A
(1)
, ··· , A
(n)
),
where d is the volume form induced by the determinant. Then we can write as
det A = d
A
(1)
, ··· ,
n
X
i=1
A
ij
e
i
, ··· , A
(n)
!
=
n
X
i=1
A
ij
d(A
(1)
, ··· , e
i
, ··· , A
(n)
)
The volume form on the right is the determinant of a matrix with the
j
th column
replaced with
e
i
. We can move our columns around so that our matrix becomes
B =
ˆ
A
ij
0
stuff 1
We get that
det B
=
det
ˆ
A
ij
, since the only permutations that give a non-zero
sum are those that send
n
to
n
. In the row and column swapping, we have made
n j column transpositions and n i row transpositions. So we have
det A =
n
X
i=1
A
ij
(1)
nj
(1)
ni
det B
=
n
X
i=1
A
ij
(1)
i+j
det
ˆ
A
ij
.
This is not only useful for computing determinants, but also computing
inverses.
Definition
(Adjugate matrix)
.
Let
A Mat
n
(
F
). The adjugate matrix of
A
,
written adj A, is the n × n matrix such that (adj A)
ij
= (1)
i+j
det
ˆ
A
ji
.
The relevance is the following result:
Theorem.
If
A Mat
n
(
F
), then
A
(
adj A
) = (
det A
)
I
n
= (
adj A
)
A
. In particu-
lar, if det A 6= 0, then
A
1
=
1
det A
adj A.
Note that this is not an efficient way to compute the inverse.
Proof. We compute
[(adj A)A]
jk
=
n
X
i=1
(adj A)
ji
A
ik
=
n
X
i=1
(1)
i+j
det
ˆ
A
ij
A
ik
. ()
So if j = k, then [(adj A)A]
jk
= det A by the lemma.
Otherwise, if
j 6
=
k
, consider the matrix
B
obtained from
A
by replacing the
j
th column by the
k
th column. Then the right hand side of (
) is just
det B
by
the lemma. But we know that if two columns are the same, the determinant is
zero. So the right hand side of () is zero. So
[(adj A)A]
jk
= det
jk
The calculation for [
A adj A
] = (
det A
)
I
n
can be done in a similar manner, or by
considering (A adj A)
T
= (adj A)
T
A
T
= (adj(A
T
))A
T
= (det A)I
n
.
Note that the coefficients of (
adj A
) are just given by polynomials in the
entries of
A
, and so is the determinant. So if
A
is invertible, then its inverse is
given by a rational function (i.e. ratio of two polynomials) in the entries of A.
This is very useful theoretically, but not computationally, since the polyno-
mials are very large. There are better ways computationally, such as Gaussian
elimination.
We’ll end with a useful tricks to compute the determinant.
Lemma. Let A, B be square matrices. Then for any C, we have
det
A C
0 B
= (det A)(det B).
Proof. Suppose A Mat
k
(F), and B Mat
`
(F), so C Mat
k,`
(F). Let
X =
A C
0 B
.
Then by definition, we have
det X =
X
σS
k+`
ε(σ)
k+`
Y
i=1
X
(i)
.
If
j k
and
i > k
, then
X
ij
= 0. We only want to sum over permutations
σ
such
that
σ
(
i
)
> k
if
i > k
. So we are permuting the last
j
things among themselves,
and hence the first
k
things among themselves. So we can decompose this into
σ
=
σ
1
σ
2
, where
σ
1
is a permutation of
{
1
, ··· , k}
and fixes the remaining things,
while σ
2
fixes {1, ··· , k}, and permutes the remaining. Then
det X =
X
σ=σ
1
σ
2
ε(σ
1
σ
2
)
k
Y
i=1
X
1
(i)
`
Y
j=1
X
k+j σ
2
(k+j)
=
X
σ
1
S
k
ε(σ
1
)
k
Y
i=1
A
1
(i)
!
X
σ
2
S
`
ε(σ
2
)
`
Y
j=1
B
jσ
2
(j)
= (det A)(det B)
Corollary.
det
A
1
stuff
A
2
.
.
.
0 A
n
=
n
Y
i=1
det A
i
6 Endomorphisms
Endomorphisms are linear maps from a vector space
V
to itself. One might
wonder why would we want to study these linear maps in particular, when
we can just work with arbitrary linear maps from any space to any other space?
When we work with arbitrary linear maps, we are free to choose any basis
for the domain, and any basis for the co-domain, since it doesn’t make sense to
require they have the “same” basis. Then we proved that by choosing the right
bases, we can put matrices into a nice form with only 1’s in the diagonal.
However, when working with endomorphisms, we can require ourselves to use
the same basis for the domain and co-domain, and there is much more we can
say. One major objective is to classify all matrices up to similarity, where two
matrices are similar if they represent the same endomorphism under different
bases.
6.1 Invariants
Definition.
If
V
is a (finite-dimensional) vector space over
F
. An endomorphism
of
V
is a linear map
α
:
V V
. We write
End
(
V
) for the
F
-vector space of all
such linear maps, and I for the identity map V V .
When we think about matrices representing an endomorphism of
V
, we’ll
use the same basis for the domain and the range. We are going to study some
properties of these endomorphisms that are not dependent on the basis we pick,
known as invariants.
Lemma.
Suppose (
e
1
, ··· , e
n
) and (
f
1
, ··· , f
n
) are bases for
V
and
α End
(
V
).
If
A
represents
α
with respect to (
e
1
, ··· , e
n
) and
B
represents
α
with respect
to (f
1
, ··· , f
n
), then
B = P
1
AP,
where P is given by
f
i
=
n
X
j=1
P
ji
e
j
.
Proof.
This is merely a special case of an earlier more general result for arbitrary
maps and spaces.
Definition
(Similar matrices)
.
We say matrices
A
and
B
are similar or conjugate
if there is some P invertible such that B = P
1
AP .
Recall that
GL
n
(
F
), the group of invertible
n × n
matrices.
GL
n
(
F
) acts on
Mat
n
(F) by conjugation:
(P, A) 7→ P AP
1
.
We are conjugating it this way so that the associativity axiom holds (otherwise
we get a right action instead of a left action). Then
A
and
B
are similar iff they
are in the same orbit. Since orbits always partition the set, this is an equivalence
relation.
Our main goal is to classify the orbits, i.e. find a “nice” representative for
each orbit.
Our initial strategy is to identify basis-independent invariants for endomor-
phisms. For example, we will show that the rank, trace, determinant and
characteristic polynomial are all such invariants.
Recall that the trace of a matrix
A Mat
n
(
F
) is the sum of the diagonal
elements:
Definition (Trace). The trace of a matrix of A Mat
n
(F) is defined by
tr A =
n
X
i=1
A
ii
.
We want to show that the trace is an invariant. In fact, we will show a
stronger statement (as well as the corresponding statement for determinants):
Lemma.
(i) If A Mat
m,n
(F) and B Mat
n,m
(F), then
tr AB = tr BA.
(ii) If A, B Mat
n
(F) are similar, then tr A = tr B.
(iii) If A, B Mat
n
(F) are similar, then det A = det B.
Proof.
(i) We have
tr AB =
m
X
i=1
(AB)
ii
=
m
X
i=1
n
X
j=1
A
ij
B
ji
=
n
X
j=1
m
X
i=1
B
ji
A
ij
= tr BA.
(ii) Suppose B = P
1
AP . Then we have
tr B = tr(P
1
(AP )) = tr((AP )P
1
) = tr A.
(iii) We have
det(P
1
AP ) = det P
1
det A det P = (det P )
1
det A det P = det A.
This allows us to define the trace and determinant of an endomorphism.
Definition
(Trace and determinant of endomorphism)
.
Let
α End
(
V
), and
A
be a matrix representing
α
under any basis. Then the trace of
α
is
tr α
=
tr A
,
and the determinant is det α = det A.
The lemma tells us that the determinant and trace are well-defined. We
can also define the determinant without reference to a basis, by defining more
general volume forms and define the determinant as a scaling factor.
The trace is slightly more tricky to define without basis, but in IB Analysis
II example sheet 4, you will find that it is the directional derivative of the
determinant at the origin.
To talk about the characteristic polynomial, we need to know what eigenvalues
are.
Definition
(Eigenvalue and eigenvector)
.
Let
α End
(
V
). Then
λ F
is an
eigenvalue (or E-value) if there is some v V \ {0} such that αv = λv.
v is an eigenvector if α(v) = λv for some λ F.
When
λ F
, the
λ
-eigenspace, written
E
α
(
λ
) or
E
(
λ
) is the subspace of
V
containing all the λ-eigenvectors, i.e.
E
α
(λ) = ker(λι α).
where ι is the identity function.
Definition
(Characteristic polynomial)
.
The characteristic polynomial of
α
is
defined by
χ
α
(t) = det( α).
You might be used to the definition
χ
α
(
t
) =
det
(
α
) instead. These two
definitions are obviously equivalent up to a factor of
1, but this definition
has an advantage that
χ
α
(
t
) is always monic, i.e. the leading coefficient is 1.
However, when doing computations in reality, we often use
det
(
α
) instead,
since it is easier to negate than α.
We know that
λ
is an eigenvalue of
α
iff
n
(
α λι
)
>
0 iff
r
(
α λι
)
< dim V
iff
χ
α
(
λ
) =
det
(
λι α
) = 0. So the eigenvalues are precisely the roots of the
characteristic polynomial.
If A Mat
n
(F), we can define χ
A
(t) = det(tI A).
Lemma.
If
A
and
B
are similar, then they have the same characteristic polyno-
mial.
Proof.
det(tI P
1
AP ) = det(P
1
(tI A)P ) = det(tI A).
Lemma. Let α End(V ) and λ
1
, ··· , λ
k
distinct eigenvalues of α. Then
E(λ
1
) + ··· + E(λ
k
) =
k
M
i=1
E(λ
i
)
is a direct sum.
Proof. Suppose
k
X
i=1
x
i
=
k
X
i=1
y
i
,
with
x
i
, y
i
E
(
λ
i
). We want to show that they are equal. We are going to find
some clever map that tells us what
x
i
and
y
i
are. Consider
β
j
End
(
V
) defined
by
β
j
=
Y
r6=j
(α λ
r
ι).
Then
β
j
k
X
i=1
x
i
!
=
k
X
i=1
Y
r6=j
(α λ
r
ι)(x
i
)
=
k
X
i=1
Y
r6=j
(λ
i
λ
r
)(x
i
).
Each summand is zero, unless i 6= j. So this is equal to
β
j
k
X
i=1
x
i
!
=
Y
r6=j
(λ
j
λ
r
)(x
j
).
Similarly, we obtain
β
j
k
X
i=1
y
i
!
=
Y
r6=j
(λ
j
λ
r
)(y
j
).
Since we know that
P
x
i
=
P
y
i
, we must have
Y
r6=j
(λ
j
λ
r
)x
j
=
Y
r6=j
(λ
j
λ
r
)y
j
.
Since we know that
Q
r6=j
(λ
r
λ
j
) 6= 0, we must have x
i
= y
i
for all i.
So each expression for
P
x
i
is unique.
The proof shows that any set of non-zero eigenvectors with distinct eigenvalues
is linearly independent.
Definition
(Diagonalizable)
.
We say
α End
(
V
) is diagonalizable if there is
some basis for
V
such that
α
is represented by a diagonal matrix, i.e. all terms
not on the diagonal are zero.
These are in some sense the nice matrices we like to work with.
Theorem.
Let
α End
(
V
) and
λ
1
, ··· , λ
k
be distinct eigenvalues of
α
. Write
E
i
for E(λ
i
). Then the following are equivalent:
(i) α is diagonalizable.
(ii) V has a basis of eigenvectors for α.
(iii) V =
L
k
i=1
E
i
.
(iv) dim V =
P
k
i=1
dim E
i
.
Proof.
(i) (ii): Suppose (e
1
, ··· , e
n
) is a basis for V . Then
α(e
i
) = A
ji
e
j
,
where
A
represents
α
. Then
A
is diagonal iff each
e
i
is an eigenvector. So
done
(ii)
(iii): It is clear that (ii) is true iff
P
E
i
=
V
, but we know that this
must be a direct sum. So done.
(iii)
(iv): This follows from example sheet 1 Q10, which says that
V
=
L
k
i=1
E
i
iff the bases for
E
i
are disjoint and their union is a basis of
V .
6.2 The minimal polynomial
6.2.1 Aside on polynomials
Before we talk about minimal polynomials, we first talk about polynomials in
general.
Definition (Polynomial). A polynomial over F is an object of the form
f(t) = a
m
t
m
+ a
m1
t
m1
+ ··· + a
1
t + a
0
,
with m 0, a
0
, ··· , a
m
F.
We write F[t] for the set of polynomials over F.
Note that we don’t identify a polynomial
f
with the corresponding function
it represents. For example, if
F
=
Z/pZ
, then
t
p
and
t
are different polynomials,
even though they define the same function (by Fermat’s little theorem/Lagrange’s
theorem). Two polynomials are equal if and only if they have the same coeffi-
cients.
However, we will later see that if
F
is
R
or
C
, then polynomials are equal
if and only if they represent the same function, and this distinction is not as
important.
Definition
(Degree)
.
Let
f F
[
t
]. Then the degree of
f
, written
deg f
is the
largest n such that a
n
6= 0. In particular, deg 0 = −∞.
Notice that deg fg = deg f + deg g and deg f + g max{deg f, deg g}.
Lemma
(Polynomial division)
.
If
f, g F
[
t
] (and
g 6
= 0), then there exists
q, r F[t] with deg r < deg g such that
f = qg + r.
Proof is omitted.
Lemma. If λ F is a root of f, i.e. f (λ) = 0, then there is some g such that
f(t) = (t λ)g(t).
Proof. By polynomial division, let
f(t) = (t λ)g(t) + r(t)
for some
g
(
t
)
, r
(
t
)
F
[
t
] with
deg r < deg
(
t λ
) = 1. So
r
has to be constant,
i.e. r(t) = a
0
for some a
0
F. Now evaluate this at λ. So
0 = f(λ) = (λ λ)g(λ) + r(λ) = a
0
.
So a
0
= 0. So r = 0. So done.
Definition
(Multiplicity of a root)
.
Let
f F
[
t
] and
λ
a root of
f
. We say
λ
has multiplicity k if (t λ)
k
is a factor of f but (t λ)
k+1
is not, i.e.
f(t) = (t λ)
k
g(t)
for some g(t) F[t] with g(λ) 6= 0.
We can use the last lemma and induction to show that any non-zero
f F
[
t
]
can be written as
f = g(t)
k
Y
i=1
(t λ
i
)
a
i
,
where
λ
1
, ··· , λ
k
are all distinct,
a
i
>
1, and
g
is a polynomial with no roots in
F.
Hence we obtain the following:
Lemma.
A non-zero polynomial
f F
[
t
] has at most
deg f
roots, counted with
multiplicity.
Corollary.
Let
f, g F
[
t
] have degree
< n
. If there are
λ
1
, ··· , λ
n
distinct such
that f (λ
i
) = g(λ
i
) for all i, then f = g.
Proof.
Given the lemma, consider
f g
. This has degree less than
n
, and
(
f g
)(
λ
i
) = 0 for
i
= 1
, ··· , n
. Since it has at least
n deg
(
f g
) roots, we
must have f g = 0. So f = g.
Corollary.
If
F
is infinite, then
f
and
g
are equal if and only if they agree on
all points.
More importantly, we have the following:
Theorem
(The fundamental theorem of algebra)
.
Every non-constant polyno-
mial over C has a root in C.
We will not prove this.
We say C is an algebraically closed field.
It thus follows that every polynomial over
C
of degree
n >
0 has precisely
n
roots, counted with multiplicity, since if we write
f
(
t
) =
g
(
t
)
Q
(
t λ
i
)
a
i
and
g
has no roots, then
g
is constant. So the number of roots is
P
a
i
=
deg f
,
counted with multiplicity.
It also follows that every polynomial in
R
factors into linear polynomials and
quadratic polynomials with no real roots (since complex roots of real polynomials
come in complex conjugate pairs).
6.2.2 Minimal polynomial
Notation.
Given
f
(
t
) =
P
m
i=0
a
i
t
i
F
[
t
],
A Mat
n
(
F
) and
α End
(
V
), we
can write
f(A) =
m
X
i=0
a
i
A
i
, f (α) =
m
X
i=0
a
i
α
i
where A
0
= I and α
0
= ι.
Theorem
(Diagonalizability theorem)
.
Suppose
α End
(
V
). Then
α
is diago-
nalizable if and only if there exists non-zero
p
(
t
)
F
[
t
] such that
p
(
α
) = 0, and
p(t) can be factored as a product of distinct linear factors.
Proof.
Suppose
α
is diagonalizable. Let
λ
1
, ··· , λ
k
be the distinct eigenvalues
of α. We have
V =
k
M
i=1
E(λ
i
).
So each v V can be written (uniquely) as
v =
k
X
i=1
v
i
with α(v
i
) = λ
i
v
i
.
Now let
p(t) =
k
Y
i=1
(t λ
i
).
Then for any v, we get
p(α)(v) =
k
X
i=1
p(α)(v
i
) =
k
X
i=1
p(λ
i
)v
i
= 0.
So p(α) = 0. By construction, p has distinct linear factors.
Conversely, suppose we have our polynomial
p(t) =
k
Y
i=1
(t λ
i
),
with
λ
1
, ··· , λ
k
F
distinct, and
p
(
α
) = 0 (we can wlog assume
p
is monic, i.e.
the leading coefficient is 1). We will show that
V =
k
X
i=1
E
α
(λ
i
).
In other words, we want to show that for all
v V
, there is some
v
i
E
α
(
λ
i
)
for i = 1, ··· , k such that v =
P
v
i
.
To find these v
i
out, we let
q
j
(t) =
Y
i6=j
t λ
i
λ
j
λ
i
.
This is a polynomial of degree k 1, and q
j
(λ
i
) = δ
ij
.
Now consider
q(t) =
k
X
i=1
q
i
(t).
We still have
deg q k
1, but
q
(
λ
i
) = 1 for any
i
. Since
q
and 1 agree on
k
points, we must have q = 1.
Let π
j
: V V be given by π
j
= q
j
(α). Then the above says that
k
X
j=1
π
j
= ι.
Hence given v V , we know that v =
P
π
j
v.
We now check that π
j
v E
α
(λ
j
). This is true since
(α λ
j
ι)π
j
v =
1
Q
i6=j
(λ
j
λ
i
)
k
Y
i=1
(α λ
ι
)(v) =
1
Q
i6=j
(λ
j
λ
i
)
p(α)(v) = 0.
So
αv
j
= λ
j
v
j
.
So done.
In the above proof, if
v E
α
(
λ
i
), then
π
j
(
v
) =
δ
ij
v
. So
π
i
is a projection
onto the E
α
(λ
i
).
Definition
(Minimal polynomial)
.
The minimal polynomial of
α End
(
V
) is
the non-zero monic polynomial M
α
(t) of least degree such that M
α
(α) = 0.
The monic requirement is just for things to look nice, since we can always
divide by the leading coefficient of a polynomial to get a monic version.
Note that if
A
represents
α
, then for all
p F
[
t
],
p
(
A
) represents
p
(
α
).
Thus
p
(
α
) is zero iff
p
(
A
) = 0. So the minimal polynomial of
α
is the minimal
polynomial of A if we define M
A
analogously.
There are two things we want to know whether the minimal polynomial
exists, and whether it is unique.
Existence is always guaranteed in finite-dimensional cases. If
dim V
=
n <
,
then
dim End
(
V
) =
n
2
. So
ι, α, α
2
, ··· , α
n
2
are linearly dependent. So there are
some λ
0
, ··· , λ
n
2
F not all zero such that
n
2
X
i=0
λ
i
α
i
= 0.
So deg M
α
n
2
. So we must have a minimal polynomial.
To show that the minimal polynomial is unique, we will prove the following
stronger characterization of the minimal polynomial:
Lemma.
Let
α End
(
V
), and
p F
[
t
]. Then
p
(
α
) = 0 if and only if
M
α
(
t
) is
a factor of p(t). In particular, M
α
is unique.
Proof.
For all such
p
, we can write
p
(
t
) =
q
(
t
)
M
α
(
t
) +
r
(
t
) for some
r
of degree
less than deg M
α
. Then
p(α) = q(α)M
α
(α) + r(α).
So if
r
(
α
) = 0 iff
p
(
α
) = 0. But
deg r < deg M
α
. By the minimality of
M
α
, we
must have r(α) = 0 iff r = 0. So p(α) = 0 iff M
α
(t) | p(t).
So if
M
1
and
M
2
are both minimal polynomials for
α
, then
M
1
| M
2
and
M
2
| M
1
. So
M
2
is just a scalar multiple of
M
1
. But since
M
1
and
M
2
are
monic, they must be equal.
Example. Let V = F
2
, and consider the following matrices:
A =
1 0
0 1
, B =
1 1
0 1
.
Consider the polynomial
p
(
t
) = (
t
1)
2
. We can compute
p
(
A
) =
p
(
B
) = 0. So
M
A
(
t
) and
M
B
(
t
) are factors of (
t
1)
2
. There aren’t many factors of (
t
1)
2
.
So the minimal polynomials are either (
t
1) or (
t
1)
2
. Since
A I
= 0 and
B I 6
= 0, the minimal polynomial of
A
is
t
1 and the minimal polynomial of
B is (t 1)
2
.
We can now re-state our diagonalizability theorem.
Theorem
(Diagonalizability theorem 2.0)
.
Let
α End
(
V
). Then
α
is diago-
nalizable if and only if M
α
(t) is a product of its distinct linear factors.
Proof. () This follows directly from the previous diagonalizability theorem.
(
) Suppose
α
is diagonalizable. Then there is some
p F
[
t
] non-zero such
that
p
(
α
) = 0 and
p
is a product of distinct linear factors. Since
M
α
divides
p
,
M
α
also has distinct linear factors.
Theorem.
Let
α, β End
(
V
) be both diagonalizable. Then
α
and
β
are
simultaneously diagonalizable (i.e. there exists a basis with respect to which
both are diagonal) if and only if αβ = βα.
This is important in quantum mechanics. This means that if two operators
do not commute, then they do not have a common eigenbasis. Hence we have
the uncertainty principle.
Proof.
(
) If there exists a basis (
e
1
, ··· , e
n
) for
V
such that
α
and
β
are repre-
sented by
A
and
B
respectively, with both diagonal, then by direct computation,
AB = BA. But AB represents αβ and BA represents βα. So αβ = βα.
(
) Suppose
αβ
=
βα
. The idea is to consider each eigenspace of
α
individu-
ally, and then diagonalize
β
in each of the eigenspaces. Since
α
is diagonalizable,
we can write
V =
k
M
i=1
E
α
(λ
i
),
where
λ
i
are the eigenvalues of
V
. We write
E
i
for
E
α
(
λ
i
). We want to show
that
β
sends
E
i
to itself, i.e.
β
(
E
i
)
E
i
. Let
v E
i
. Then we want
β
(
v
) to be
in E
i
. This is true since
α(β(v)) = β(α(v)) = β(λ
i
v) = λ
i
β(v).
So β(v) is an eigenvector of α with eigenvalue λ
i
.
Now we can view β|
E
i
End(E
i
). Note that
M
β
(β|
E
i
) = M
β
(β)|
E
i
= 0.
Since
M
β
(
t
) is a product of its distinct linear factors, it follows that
β|
E
i
is
diagonalizable. So we can choose a basis
B
i
of eigenvectors for
β|
E
i
. We can do
this for all i.
Then since
V
is a direct sum of the
E
i
’s, we know that
B
=
S
k
i=1
B
i
is a
basis for V consisting of eigenvectors for both α and β. So done.
6.3 The Cayley-Hamilton theorem
We will first state the theorem, and then prove it later.
Recall that
χ
α
(
t
) =
det
(
α
) for
α End
(
V
). Our main theorem of the
section (as you might have guessed from the title) is
Theorem
(Cayley-Hamilton theorem)
.
Let
V
be a finite-dimensional vector
space and
α End
(
V
). Then
χ
α
(
α
) = 0, i.e.
M
α
(
t
)
| χ
α
(
t
). In particular,
deg M
α
n.
We will not prove this yet, but just talk about it first. It is tempting to prove
this by substituting
t
=
α
into
det
(
α
) and get
det
(
α α
) = 0, but this is
meaningless, since what the statement
χ
α
(
t
) =
det
(
α
) tells us to do is to
expand the determinant of the matrix
t a
11
a
12
··· a
1n
a
21
t a
22
··· a
2n
.
.
.
.
.
.
.
.
.
.
.
.
a
n1
a
n2
··· t a
nn
to obtain a polynomial, and we clearly cannot substitute
t
=
A
in this expression.
However, we can later show that we can use this idea to prove it, but just be a
bit more careful.
Note also that if ρ(t) F[t] and
A =
λ
1
.
.
.
λ
n
,
then
ρ(A) =
ρ(λ
1
)
.
.
.
ρ(λ
n
)
.
Since
χ
A
(
t
) is defined as
Q
n
i=1
(
t λ
i
), it follows that
χ
A
(
A
) = 0. So if
α
is
diagonalizable, then the theorem is clear.
This was easy. Diagonalizable matrices are nice. The next best thing we can
look at is upper-triangular matrices.
Definition
(Triangulable)
.
An endomorphism
α End
(
V
) is triangulable if
there is a basis for V such that α is represented by an upper triangular matrix
a
11
a
12
··· a
1n
0 a
22
··· a
2n
.
.
.
.
.
.
.
.
.
.
.
.
0 0 ··· a
nn
.
We have a similar lemma telling us when matrices are triangulable.
Lemma.
An endomorphism
α
is triangulable if and only if
χ
α
(
t
) can be written
as a product of linear factors, not necessarily distinct. In particular, if
F
=
C
(or any algebraically closed field), then every endomorphism is triangulable.
Proof. Suppose that α is triangulable and represented by
λ
1
···
0 λ
2
···
.
.
.
.
.
.
.
.
.
.
.
.
0 0 ··· λ
n
.
Then
χ
α
(t) = det
t λ
1
···
0 t λ
2
···
.
.
.
.
.
.
.
.
.
.
.
.
0 0 ··· t λ
n
=
n
Y
i=1
(t λ
i
).
So it is a product of linear factors.
We are going to prove the converse by induction on the dimension of our
space. The base case
dim V
= 1 is trivial, since every 1
×
1 matrix is already
upper triangular.
Suppose
α End
(
V
) and the result holds for all spaces of dimensions
< dim V
, and
χ
α
is a product of linear factors. In particular,
χ
α
(
t
) has a root,
say λ F.
Now let
U
=
E
(
λ
)
6
= 0, and let
W
be a complementary subspace to
U
in
V
,
i.e.
V
=
U W
. Let
u
1
, ··· , u
r
be a basis for
U
and
w
r+1
, ··· , w
n
be a basis
for
W
so that
u
1
, ··· , u
r
, w
r+1
, ··· , w
n
is a basis for
V
, and
α
is represented by
λI
r
stuff
0 B
We know
χ
α
(
t
) = (
t λ
)
r
χ
B
(
t
). So
χ
B
(
t
) is also a product of linear factors. We
let β : W W be the map defined by B with respect to w
r+1
, ··· , w
n
.
(Note that in general,
β
is not
α|
W
in general, since
α
does not necessarily
map
W
to
W
. However, we can say that (
αβ
)(
w
)
U
for all
w W
. This can
be much more elegantly expressed in terms of quotient spaces, but unfortunately
that is not officially part of the course)
Since
dim W < dim V
, there is a basis
v
r+1
, ··· , v
n
for
W
such that
β
is
represented by C, which is upper triangular.
For j = 1, ··· , n r, we have
α(v
j+r
) = u +
nr
X
k=1
C
kj
v
k+r
for some u U. So α is represented by
λI
r
stuff
0 C
with respect to (u
1
, ··· , u
r
, v
r+1
, ··· , v
n
), which is upper triangular.
Example. Consider the real rotation matrix
cos θ sin θ
sin θ cos θ
.
This is not similar to a real upper triangular matrix (if
θ
is not an integer
multiple of
π
). This is since the eigenvalues are
e
±
and are not real. On the
other hand, as a complex matrix, it is triangulable, and in fact diagonalizable
since the eigenvalues are distinct.
For this reason, in the rest of the section, we are mostly going to work in
C
.
We can now prove the Cayley-Hamilton theorem.
Theorem
(Cayley-Hamilton theorem)
.
Let
V
be a finite-dimensional vector
space and
α End
(
V
). Then
χ
α
(
α
) = 0, i.e.
M
α
(
t
)
| χ
α
(
t
). In particular,
deg M
α
n.
Proof.
In this proof, we will work over
C
. By the lemma, we can choose a basis
{e
1
, ··· , e
n
} is represented by an upper triangular matrix.
A =
λ
1
···
0 λ
2
···
.
.
.
.
.
.
.
.
.
.
.
.
0 0 ··· λ
n
.
We must prove that
χ
α
(α) = χ
A
(α) =
n
Y
i=1
(α λ
i
ι) = 0.
Write V
j
= he
1
, ··· , e
j
i. So we have the inclusions
V
0
= 0 V
1
··· V
n1
V
n
= V.
We also know that dim V
j
= j. This increasing sequence is known as a flag.
Now note that since A is upper-triangular, we get
α(e
i
) =
i
X
k=1
A
ki
e
k
V
i
.
So α(V
j
) V
j
for all j = 0, ··· , n.
Moreover, we have
(α λ
j
ι)(e
j
) =
j1
X
k=1
A
kj
e
k
V
j1
for all
j
= 1
, ··· , n
. So every time we apply one of these things, we get to a
smaller space. Hence by induction on n j, we have
n
Y
i=j
(α λ
i
ι)(V
n
) V
j1
.
In particular, when j = 1, we get
n
Y
i=1
(α λ
i
ι)(V ) V
0
= 0.
So χ
α
(α) = 0 as required.
Note that if our field
F
is not
C
but just a subfield of
C
, say
R
, we can just
pretend it is a complex matrix, do the same proof.
We can see this proof more “visually” as follows: for simplicity of expression,
we suppose
n
= 4. In the basis where
α
is upper-triangular, the matrices
A λ
i
I
look like this
A λ
1
I =
0 ∗ ∗ ∗
0 ∗ ∗ ∗
0 0 ∗ ∗
0 0 0
A λ
2
I =
∗ ∗ ∗ ∗
0 0 ∗ ∗
0 0 ∗ ∗
0 0 0
A λ
3
I =
∗ ∗ ∗ ∗
0 ∗ ∗ ∗
0 0 0
0 0 0
A λ
4
I =
∗ ∗ ∗ ∗
0 ∗ ∗ ∗
0 0 ∗ ∗
0 0 0 0
Then we just multiply out directly (from the right):
4
Y
i=1
(A λ
i
I) =
0 ∗ ∗ ∗
0 ∗ ∗ ∗
0 0 ∗ ∗
0 0 0
∗ ∗ ∗ ∗
0 0 ∗ ∗
0 0 ∗ ∗
0 0 0
∗ ∗ ∗ ∗
0 ∗ ∗ ∗
0 0 0
0 0 0
∗ ∗ ∗ ∗
0 ∗ ∗ ∗
0 0 ∗ ∗
0 0 0 0
=
0 ∗ ∗ ∗
0 ∗ ∗ ∗
0 0 ∗ ∗
0 0 0
∗ ∗ ∗ ∗
0 0 ∗ ∗
0 0 ∗ ∗
0 0 0
∗ ∗ ∗ ∗
0 ∗ ∗ ∗
0 0 0 0
0 0 0 0
=
0 ∗ ∗ ∗
0 ∗ ∗ ∗
0 0 ∗ ∗
0 0 0
∗ ∗ ∗ ∗
0 0 0 0
0 0 0 0
0 0 0 0
=
0 0 0 0
0 0 0 0
0 0 0 0
0 0 0 0
.
This is exactly what we showed in the proof after multiplying out the first
k
elements of the product (counting from the right), the image is contained in the
span of the first n k basis vectors.
Proof.
We’ll now prove the theorem again, which is somewhat a formalization
of the “nonsense proof” where we just substitute t = α into det(α ).
Let α be represented by A, and B = tI A. Then
B adj B = det BI
n
= χ
α
(t)I
n
.
But we know that
adj B
is a matrix with entries in
F
[
t
] of degree at most
n
1.
So we can write
adj B = B
n1
t
n1
+ B
n2
t
n2
+ ··· + B
0
,
with B
i
Mat
n
(F). We can also write
χ
α
(t) = t
n
+ a
n1
t
n1
+ ··· + a
0
.
Then we get the result
(tI
n
A)(B
n1
t
n1
+ B
n2
t
n2
+ ··· + B
0
) = (t
n
+ a
n1
t
n1
+ ··· + a
0
)I
n
.
We would like to just throw in
t
=
A
, and get the desired result, but in all these
derivations, t is assumed to be a real number, and, tI
n
A is the matrix
t a
11
a
12
··· a
1n
a
21
t a
22
··· a
2n
.
.
.
.
.
.
.
.
.
.
.
.
a
n1
a
n2
··· t a
nn
It doesn’t make sense to put our A in there.
However, what we can do is to note that since this is true for all values of
t
,
the coefficients on both sides must be equal. Equating coefficients in
t
k
, we have
AB
0
= a
0
I
n
B
0
AB
1
= a
1
I
n
.
.
.
B
n2
AB
n1
= a
n1
I
n
AB
n1
0 = I
n
We now multiply each row a suitable power of A to obtain
AB
0
= a
0
I
n
AB
0
A
2
B
1
= a
1
A
.
.
.
A
n1
B
n2
A
n
B
n1
= a
n1
A
n1
A
n
B
n1
0 = A
n
.
Summing this up then gives χ
α
(A) = 0.
This proof suggests that we really ought to be able to just substitute in
t
=
α
and be done. In fact, we can do this, after we develop sufficient machinery. This
will be done in the IB Groups, Rings and Modules course.
Lemma. Let α End(V ), λ F. Then the following are equivalent:
(i) λ is an eigenvalue of α.
(ii) λ is a root of χ
α
(t).
(iii) λ is a root of M
α
(t).
Proof.
(i)
(ii):
λ
is an eigenvalue of
α
if and only if (
α λι
)(
v
) = 0 has a
non-trivial root, iff det(α λι) = 0.
(iii) (ii): This follows from Cayley-Hamilton theorem since M
α
| χ
α
.
(i)
(iii): Let
λ
be an eigenvalue, and
v
be a corresponding eigenvector.
Then by definition of M
α
, we have
M
α
(α)(v) = 0(v) = 0.
We also know that
M
α
(α)(v) = M
α
(λ)v.
Since v is non-zero, we must have M
α
(λ) = 0.
(iii)
(i): This is not necessary since it follows from the above, but
we could as well do it explicitly. Suppose
λ
is a root of
M
α
(
t
). Then
M
α
(
t
) = (
t λ
)
g
(
t
) for some
g F
[
t
]. But
deg g < deg M
α
. Hence by
minimality of
M
α
, we must have
g
(
α
)
6
= 0. So there is some
v V
such
that g(α)(v) 6= 0. Then
(α λι)g(α)(v) = M
α
(α)v = 0.
So we must have
α
(
g
(
α
)(
v
)) =
λg
(
α
)(
v
). So
g
(
α
)(
v
)
E
α
(
λ
)
\{
0
}
. So (i)
holds.
Example. What is the minimal polynomial of
A =
1 0 2
0 1 1
0 0 2
?
We can compute
χ
A
(
t
) = (
t
1)
2
(
t
2). So we know that the minimal polynomial
is one of (t 1)
2
(t 2) and (t 1)(t 2).
By direct and boring computations, we can find (
A I
)(
A
2
I
) = 0. So we
know that M
A
(t) = (t 1)(t 2). So A is diagonalizable.
6.4 Multiplicities of eigenvalues and Jordan normal form
We will want to put our matrices in their “Jordan normal forms”, which is
a unique form for each equivalence class of similar matrices. The following
properties will help determine which Jordan normal form a matrix can have.
Definition
(Algebraic and geometry multiplicity)
.
Let
α End
(
V
) and
λ
an
eigenvalue of α. Then
(i)
The algebraic multiplicity of
λ
, written
a
λ
, is the multiplicity of
λ
as a
root of χ
α
(t).
(ii)
The geometric multiplicity of
λ
, written
g
λ
, is the dimension of the corre-
sponding eigenspace, dim E
α
(λ).
(iii) c
λ
is the multiplicity of λ as a root of the minimal polynomial m
α
(t).
We will look at some extreme examples:
Example.
Let
A =
λ 1 ··· 0
0 λ
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
1
0 0 ··· λ
.
We will later show that a
λ
= n = c
λ
and g
λ
= 1.
Consider A = λI. Then a
λ
= g
λ
= n, c
λ
= 1.
Lemma. If λ is an eigenvalue of α, then
(i) 1 g
λ
a
λ
(ii) 1 c
λ
a
λ
.
Proof.
(i)
The first inequality is easy. If
λ
is an eigenvalue, then
E
(
λ
)
6
= 0. So
g
λ
=
dim E
(
λ
)
1. To prove the other inequality, if
v
1
, ··· , v
g
is a basis
for
E
(
λ
), then we can extend it to a basis for
V
, and then
α
is represented
by
λI
g
0 B
So χ
α
(t) = (t λ)
g
χ
B
(t). So a
λ
> g = g
λ
.
(ii)
This is straightforward since
M
α
(
λ
) = 0 implies 1
c
λ
, and since
M
α
(
t
)
|
χ
α
(t), we know that c
λ
α
λ
.
Lemma. Suppose F = C and α End(V ). Then the following are equivalent:
(i) α is diagonalizable.
(ii) g
λ
= a
λ
for all eigenvalues of α.
(iii) c
λ
= 1 for all λ.
Proof.
(i)
(ii):
α
is diagonalizable iff
dim V
=
P
dim E
α
(
λ
i
). But this is
equivalent to
dim V =
X
g
λ
i
X
a
λ
i
= deg χ
α
= dim V.
So we must have
P
g
λ
i
=
P
a
λ
i
. Since each
g
λ
i
is at most
a
λ
i
, they must
be individually equal.
(i)
(iii):
α
is diagonalizable if and only if
M
α
(
t
) is a product of distinct
linear factors if and only if c
λ
= 1 for all eigenvalues λ.
Definition
(Jordan normal form)
.
We say
A Mat
N
(
C
) is in Jordan normal
form if it is a block diagonal of the form
J
n
1
(λ
1
) 0
J
n
2
(λ
2
)
.
.
.
0 J
n
k
(λ
k
)
where
k
1,
n
1
, ··· , n
k
N
such that
n
=
P
n
i
,
λ
1
, ··· , λ
k
not necessarily
distinct, and
J
m
(λ) =
λ 1 ··· 0
0 λ
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
1
0 0 ··· λ
is an m × m matrix. Note that J
m
(λ) = λI
m
+ J
m
(0).
For example, it might look something like
λ
1
1
λ
1
1
λ
1
0
λ
2
0
λ
3
1
λ
3
0
.
.
.
.
.
.
λ
n
1
λ
n
with all other entries zero.
Then we have the following theorem:
Theorem
(Jordan normal form theorem)
.
Every matrix
A Mat
n
(
C
) is similar
to a matrix in Jordan normal form. Moreover, this Jordan normal form matrix
is unique up to permutation of the blocks.
This is a complete solution to the classification problem of matrices, at least
in
C
. We will not prove this completely. We will only prove the uniqueness part,
and then reduce the existence part to a special form of endomorphisms. The
remainder of the proof is left for IB Groups, Rings and Modules.
We can rephrase this result using linear maps. If
α End
(
V
) is an endo-
morphism of a finite-dimensional vector space
V
over
C
, then the theorem says
there exists a basis such that
α
is represented by a matrix in Jordan normal
form, and this is unique as before.
Note that the permutation thing is necessary, since if two matrices in Jordan
normal form differ only by a rearrangement of blocks, then they are similar, by
permuting the basis.
Example.
Every 2
×
2 matrix in Jordan normal form is one of the three types:
Jordan normal form χ
A
M
A
λ 0
0 λ
(t λ)
2
(t λ)
λ 0
0 µ
(t λ)(t µ) (t λ)(t µ)
λ 1
0 λ
(t λ)
2
(t λ)
2
with
λ, µ C
distinct. We see that
M
A
determines the Jordan normal form of
A, but χ
A
does not.
Every 3
×
3 matrix in Jordan normal form is one of the six types. Here
λ
1
, λ
2
and λ
3
are distinct complex numbers.
Jordan normal form χ
A
M
A
λ
1
0 0
0 λ
2
0
0 0 λ
3
(t λ
1
)(t λ
2
)(t λ
3
) (t λ
1
)(t λ
2
)(t λ
3
)
λ
1
0 0
0 λ
1
0
0 0 λ
2
(t λ
1
)
2
(t λ
2
) (t λ
1
)(t λ
2
)
λ
1
1 0
0 λ
1
0
0 0 λ
2
(t λ
1
)
2
(t λ
2
) (t λ
1
)
2
(t λ
2
)
λ
1
0 0
0 λ
1
0
0 0 λ
1
(t λ
1
)
3
(t λ
1
)
λ
1
1 0
0 λ
1
0
0 0 λ
1
(t λ
1
)
3
(t λ
1
)
2
λ
1
1 0
0 λ
1
1
0 0 λ
1
(t λ
1
)
3
(t λ
1
)
3
Notice that
χ
A
and
M
A
together determine the Jordan normal form of a 3
×
3
complex matrix. We do indeed need
χ
A
in the second case, since if we are given
M
A
= (t λ
1
)(t λ
2
), we know one of the roots is double, but not which one.
In general, though, even χ
A
and M
A
together does not suffice.
We now want to understand the Jordan normal blocks better. Recall the
definition
J
n
(λ) =
λ 1 ··· 0
0 λ
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
1
0 0 ··· λ
= λI
n
+ J
n
(0).
If (e
1
, ··· , e
n
) is the standard basis for C
n
, we have
J
n
(0)(e
1
) = 0, J
n
(0)(e
i
) = e
i1
for 2 i n.
Thus we know
J
n
(0)
k
(e
i
) =
(
0 i k
e
ik
k < i n
In other words, for k < n, we have
(J
n
(λ) λI)
k
= J
n
(0)
k
=
0 I
nk
0 0
.
If k n, then we have (J
n
(λ) λI)
k
= 0. Hence we can summarize this as
n((J
m
(λ) λI
m
)
r
) = min{r, m}.
Note that if
A
=
J
n
(
λ
), then
χ
A
(
t
) =
M
A
(
t
) = (
t λ
)
n
. So
λ
is the only
eigenvalue of
A
. So
a
λ
=
c
λ
=
n
. We also know that
n
(
AλI
) =
nr
(
AλI
) =
1. So g
λ
= 1.
Recall that a general Jordan normal form is a block diagonal matrix of Jordan
blocks. We have just studied individual Jordan blocks. Next, we want to look at
some properties of block diagonal matrices in general. If
A
is the block diagonal
matrix
A =
A
1
A
2
.
.
.
A
k
,
then
χ
A
(t) =
k
Y
i=1
χ
A
i
(k).
Moreover, if ρ F[t], then
ρ(A) =
ρ(A
1
)
ρ(A
2
)
.
.
.
ρ(A
k
)
.
Hence
M
A
(t) = lcm(M
A
1
(t), ··· , M
A
k
(t)).
By the rank-nullity theorem, we have
n(ρ(A)) =
k
X
i=1
n(ρ(A
i
)).
Thus if A is in Jordan normal form, we get the following:
g
λ
is the number of Jordan blocks in A with eigenvalue λ.
a
λ
is the sum of sizes of the Jordan blocks of A with eigenvalue λ.
c
λ
is the size of the largest Jordan block with eigenvalue λ.
We are now going to prove the uniqueness part of Jordan normal form
theorem.
Theorem.
Let
α End
(
V
), and
A
in Jordan normal form representing
α
. Then
the number of Jordan blocks J
n
(λ) in A with n r is
n((α λι)
r
) n((α λι)
r1
).
This is clearly independent of the choice of basis. Also, given this information,
we can figure out how many Jordan blocks of size exactly
n
by doing the right
subtraction. Hence this tells us that Jordan normal forms are unique up to
permutation of blocks.
Proof. We work blockwise for
A =
J
n
1
(λ
1
)
J
n
2
(λ
2
)
.
.
.
J
n
k
(λ
k
)
.
We have previously computed
n((J
m
(λ) λI
m
)
r
) =
(
r r m
m r > m
.
Hence we know
n((J
m
(λ) λI
m
)
r
) n((J
m
(λ) λI
m
)
r1
) =
(
1 r m
0 otherwise.
It is also easy to see that for µ 6= λ,
n((J
m
(µ) λI
m
)
r
) = n(J
m
(µ λ)
r
) = 0
Adding up for each block, for r 1, we have
n((αλι)
r
)n((αλι)
r1
) = number of Jordan blocks J
n
(λ) with n r.
We can interpret this result as follows: if
r m
, when we take an additional
power of
J
m
(
λ
)
λI
m
, we get from
0 I
mr
0 0
to
0 I
mr1
0 0
. So we kill off
one more column in the matrix, and the nullity increase by one. This happens
until (
J
m
(
λ
)
λI
m
)
r
= 0, in which case increasing the power no longer affects
the matrix. So when we look at the difference in nullity, we are counting the
number of blocks that are affected by the increase in power, which is the number
of blocks of size at least r.
We have now proved uniqueness, but existence is not yet clear. To show
this, we will reduce it to the case where there is exactly one eigenvalue. This
reduction is easy if the matrix is diagonalizable, because we can decompose the
matrix into each eigenspace and then work in the corresponding eigenspace. In
general, we need to work with “generalized eigenspaces”.
Theorem
(Generalized eigenspace decomposition)
.
Let
V
be a finite-dimensional
vector space C such that α End(V ). Suppose that
M
α
(t) =
k
Y
i=1
(t λ
i
)
c
i
,
with λ
1
, ··· , λ
k
C distinct. Then
V = V
1
··· V
k
,
where V
i
= ker((α λ
i
ι)
c
i
) is the generalized eigenspace.
This allows us to decompose
V
into a block diagonal matrix, and then each
block will only have one eigenvalue.
Note that if
c
1
=
···
=
c
k
= 1, then we recover the diagonalizability theorem.
Hence, it is not surprising that the proof of this is similar to the diagonalizability
theorem. We will again prove this by constructing projection maps to each of
the V
i
.
Proof. Let
p
j
(t) =
Y
i6=j
(t λ
i
)
c
i
.
Then
p
1
, ··· , p
k
have no common factors, i.e. they are coprime. Thus by Euclid’s
algorithm, there exists q
1
, ··· , q
k
C[t] such that
X
p
i
q
i
= 1.
We now define the endomorphism
π
j
= q
j
(α)p
j
(α)
for j = 1, ··· , k.
Then
P
π
j
= ι. Since M
α
(α) = 0 and M
α
(t) = (t λ
j
ι)
c
j
p
j
(t), we get
(α λ
j
ι)
c
j
π
j
= 0.
So im π
j
V
j
.
Now suppose v V . Then
v = ι(v) =
k
X
j=1
π
j
(v)
X
V
j
.
So
V =
X
V
j
.
To show this is a direct sum, note that
π
i
π
j
= 0, since the product contains
M
α
(α) as a factor. So
π
i
= ιπ
i
=
X
π
j
π
i
= π
2
i
.
So π is a projection, and π
j
|
V
j
= ι
V
j
. So if v =
P
v
i
, then applying π
i
to both
sides gives
v
i
=
π
i
(
v
). Hence there is a unique way of writing
v
as a sum of
things in V
i
. So V =
L
V
j
as claimed.
Note that we didn’t really use the fact that the vector space is over
C
, except
to get that the minimal polynomial is a product of linear factors. In fact, for
arbitrary vector spaces, if the minimal polynomial of a matrix is a product of
linear factors, then it can be put into Jordan normal form. The converse is also
true if it can be put into Jordan normal form, then the minimal polynomial
is a product of linear factors, since we’ve seen that a necessary and sufficient
condition for the minimal polynomial to be a product of linear factors is for
there to be a basis in which the matrix is upper triangular.
Using this theorem, by restricting
α
to its generalized eigenspaces, we can
reduce the existence part of the Jordan normal form theorem to the case
M
α
(
t
) =
(
t λ
)
c
. Further by replacing
α
by
α λι
, we can reduce this to the case where
0 is the only eigenvalue.
Definition
(Nilpotent)
.
We say
α End
(
V
) is nilpotent if there is some
r
such
that α
r
= 0.
Over
C
,
α
is nilpotent if and only if the only eigenvalue of
α
is 0. This is
since α is nilpotent if the minimal polynomial is t
r
for some r.
We’ve now reduced the problem of classifying complex endomorphisms to the
classifying nilpotent endomorphisms. This is the point where we stop. For the
remainder of the proof, see IB Groups, Rings and Modules. There is in fact an
elementary proof of it, and we’re not doing it not because it’s hard, but because
we don’t have time.
Example. Let
A =
3 2 0
1 0 0
1 0 1
We know we can find the Jordan normal form by just computing the minimal
polynomial and characteristic polynomial. But we can do better and try to find
a P such that P
1
AP is in Jordan normal form.
We first compute the eigenvalues of A. The characteristic polynomial is
det
t 3 2 0
1 t 0
1 0 t 1
= (t 1)((t 3)t + 2) = (t 1)
2
(t 2).
We now compute the eigenspaces of A. We have
A I =
2 2 0
1 1 0
1 0 0
We see this has rank 2 and hence nullity 1, and the eigenspace is the kernel
E
A
(1) =
*
0
0
1
+
We can also compute the other eigenspace. We have
A 2I =
1 2 0
1 2 0
1 0 1
This has rank 2 and
E
A
(2) =
*
2
1
2
+
.
Since
dim E
A
(1) + dim E
A
(2) = 2 < 3,
this is not diagonalizable. So the minimal polynomial must also be
M
A
(
t
) =
χ
A
(
t
) = (
t
1)
2
(
t
2). From the classificaion last time, we know that
A
is
similar to
1 1 0
0 1 0
0 0 2
We now want to compute a basis that transforms
A
to this. We want a basis
(v
1
, v
2
, v
3
) of C
3
such that
Av
1
= v
1
, Av
2
= v
1
+ v
2
, Av
3
= 2v
3
.
Equivalently, we have
(A I)v
1
= 0, (A I)v
2
= v
1
, (A 2I)v
3
= 0.
There is an obvious choices v
3
, namely the eigenvector of eigenvalue 2.
To find
v
1
and
v
2
, the idea is to find some
v
2
such that (
A I
)
v
2
6
=
0
but
(A I)
2
v
2
= 0. Then we can let v
1
= (A I)v
1
.
We can compute the kernel of (A I)
2
. We have
(A I)
2
=
2 2 0
1 1 0
2 2 0
The kernel of this is
ker(A I)
2
=
*
0
0
1
,
1
1
0
+
.
We need to pick our
v
2
that is in this kernel but not in the kernel of
A I
(which is the eigenspace E
1
we have computed above). So we have
v
2
=
1
1
0
, v
1
=
0
0
1
, v
3
=
2
1
2
.
Hence we have
P =
0 1 2
0 1 1
1 0 2
and
P
1
AP =
1 1 0
0 1 0
0 0 2
.
7 Bilinear forms II
In Chapter 4, we have looked at bilinear forms in general. Here, we want to look
at bilinear forms on a single space, since often there is just one space we are
interested in. We are also not looking into general bilinear forms on a single
space, but just those that are symmetric.
7.1 Symmetric bilinear forms and quadratic forms
Definition
(Symmetric bilinear form)
.
Let
V
is a vector space over
F
. A bilinear
form φ : V ×V F is symmetric if
φ(v, w) = φ(w, v)
for all v, w V .
Example.
If
S Mat
n
(
F
) is a symmetric matrix, i.e.
S
T
=
S
, the bilinear form
φ : F
n
× F
n
F defined by
φ(x, y) = x
T
Sx =
n
X
i,j=1
x
i
S
ij
y
j
is a symmetric bilinear form.
This example is typical in the following sense:
Lemma.
Let
V
be a finite-dimensional vector space over
F
, and
φ
:
V ×V F
is a symmetric bilinear form. Let (
e
1
, ··· , e
n
) be a basis for
V
, and let
M
be
the matrix representing
φ
with respect to this basis, i.e.
M
ij
=
φ
(
e
i
, e
j
). Then
φ is symmetric if and only if M is symmetric.
Proof. If φ is symmetric, then
M
ij
= φ(e
i
, e
j
) = φ(e
j
, e
i
) = M
ji
.
So M
T
= M. So M is symmetric.
If M is symmetric, then
φ(x, y) = φ
X
x
i
e
i
,
X
y
j
e
j
=
X
i,j
x
i
M
ij
y
j
=
n
X
i,j
y
j
M
ji
x
i
= φ(y, x).
We are going to see what happens when we change basis. As in the case of
endomorphisms, we will require to change basis in the same ways on both sides.
Lemma.
Let
V
is a finite-dimensional vector space, and
φ
:
V × V F
a
bilinear form. Let (e
1
, ··· , e
n
) and (f
1
, ··· , f
n
) be bases of V such that
f
i
=
n
X
k=1
P
ki
e
k
.
If
A
represents
φ
with respect to (
e
i
) and
B
represents
φ
with respect to (
f
i
),
then
B = P
T
AP.
Proof. Special case of general formula proven before.
This motivates the following definition:
Definition
(Congruent matrices)
.
Two square matrices
A, B
are congruent if
there exists some invertible P such that
B = P
T
AP.
It is easy to see that congruence is an equivalence relation. Two matrices
are congruent if and only if represent the same bilinear form with respect to
different bases.
Thus, to classify (symmetric) bilinear forms is the same as classifying (sym-
metric) matrices up to congruence.
Before we do the classification, we first look at quadratic forms, which are
something derived from bilinear forms.
Definition
(Quadratic form)
.
A function
q
:
V F
is a quadratic form if there
exists some bilinear form φ such that
q(v) = φ(v, v)
for all v V .
Note that quadratic forms are not linear maps (they are quadratic).
Example.
Let
V
=
R
2
and
φ
be represented by
A
with respect to the standard
basis. Then
q

x
y

=
x y
A
11
A
12
A
21
A
22
x
y
= A
11
x
2
+ (A
12
+ A
21
)xy + A
22
y
2
.
Notice that if A is replaced the symmetric matrix
1
2
(A + A
T
),
then we get a different φ, but the same q. This is in fact true in general.
Proposition
(Polarization identity)
.
Suppose that
char F 6
= 2, i.e. 1 + 1
6
= 0
on
F
(e.g. if
F
is
R
or
C
). If
q
:
V F
is a quadratic form, then there exists a
unique symmetric bilinear form φ : V × V F such that
q(v) = φ(v, v).
Proof.
Let
ψ
:
V × V F
be a bilinear form such that
ψ
(
v, v
) =
q
(
v
). We
define φ : V ×V F by
φ(v, w) =
1
2
(ψ(v, w) + ψ(w, v))
for all
v, w F
. This is clearly a bilinear form, and it is also clearly symmetric
and satisfies the condition we wants. So we have proved the existence part.
To prove uniqueness, we want to find out the values of
φ
(
v, w
) in terms of
what q tells us. Suppose φ is such a symmetric bilinear form. We compute
q(v + w) = φ(v + w, v + w)
= φ(v, v) + φ(v, w) + φ(w, v) + φ(w, w)
= q(v) + 2φ(v, w) + q(w).
So we have
φ(v, w) =
1
2
(q(v + w) q(v) q(w)).
So it is determined by q, and hence unique.
Theorem.
Let
V
be a finite-dimensional vector space over
F
, and
φ
:
V ×V F
a symmetric bilinear form. Then there exists a basis (
e
1
, ··· , e
n
) for
V
such
that φ is represented by a diagonal matrix with respect to this basis.
This tells us classifying symmetric bilinear forms is easier than classifying
endomorphisms, since for endomorphisms, even over
C
, we cannot always make
it diagonal, but we can for bilinear forms over arbitrary fields.
Proof.
We induct over
n
=
dim V
. The cases
n
= 0 and
n
= 1 are trivial, since
all matrices are diagonal.
Suppose we have proven the result for all spaces of dimension less than
n
.
First consider the case where
φ
(
v, v
) = 0 for all
v V
. We want to show that
we must have
φ
= 0. This follows from the polarization identity, since this
φ
induces the zero quadratic form, and we know that there is a unique bilinear
form that induces the zero quadratic form. Since we know that the zero bilinear
form also induces the zero quadratic form, we must have
φ
= 0. Then
φ
will
be represented by the zero matrix with respect to any basis, which is trivially
diagonal.
If not, pick e
1
V such that φ(e
1
, e
1
) 6= 0. Let
U = ker φ(e
1
, ·) = {u V : φ(e
1
, u) = 0}.
Since
φ
(
e
1
, ·
)
V
\ {
0
}
, we know that
dim U
=
n
1 by the rank-nullity
theorem.
Our objective is to find other basis elements
e
2
, ··· , e
n
such that
φ
(
e
1
, e
j
) = 0
for all j > 1. For this to happen, we need to find them inside U.
Now consider
φ|
U×U
:
U × U F
, a symmetric bilinear form. By the
induction hypothesis, we can find a basis
e
2
, ··· , e
n
for
U
such that
φ|
U×U
is
represented by a diagonal matrix with respect to this basis.
Now by construction,
φ
(
e
i
, e
j
) = 0 for all 1
i 6
=
j n
and (
e
1
, ··· , e
n
) is
a basis for V . So we’re done.
Example. Let q be a quadratic form on R
3
given by
q
x
y
z
= x
2
+ y
2
+ z
2
+ 2xy + 4yz + 6xz.
We want to find a basis f
1
, f
2
, f
3
for R
3
such that q is of the form
q(af
1
+ bf
2
+ cf
3
) = λa
2
+ µb
2
+ νc
2
for some λ, µ, ν R.
There are two ways to do this. The first way is to follow the proof we just had.
We first find our symmetric bilinear form. It is the bilinear form represented by
the matrix
A =
1 1 3
1 1 2
3 2 1
.
We then find
f
1
such that
φ
(
f
1
, f
1
)
6
= 0. We note that
q
(
e
1
) = 1
6
= 0. So we pick
f
1
= e
1
=
1
0
0
.
Then
φ(e
1
, v) =
1 0 0
1 1 3
1 1 2
3 2 1
v
1
v
2
v
3
= v
1
+ v
2
+ 3v
3
.
Next we need to pick our
f
2
. Since it is in the kernel of
φ
(
f
1
, ·
), it must satisfy
φ(f
1
, f
2
) = 0.
To continue our proof inductively, we also have to pick an f
2
such that
φ(f
2
, f
2
) 6= 0.
For example, we can pick
f
2
=
3
0
1
.
Then we have q(f
2
) = 8.
Then we have
φ(f
2
, v) =
3 0 1
1 1 3
1 1 2
3 2 1
v
1
v
2
v
3
= v
2
+ 8v
3
Finally, we want φ(f
1
, f
3
) = φ(f
2
, f
3
) = 0. Then
f
3
=
5
8
1
works. We have q(f
3
) = 8.
With these basis elements, we have
q(af
1
+ bf
2
+ cf
3
) = φ(af
1
+ bf
2
+ cf
3
, af
1
+ bf
2
+ cf
3
)
= a
2
q(f
1
) + b
2
q(f
2
) + c
2
q(f
3
)
= a
2
8b
2
+ 8c
2
.
Alternatively, we can solve the problem by completing the square. We have
x
2
+ y
2
+ z
2
+ 2xy + 4yz + 6xz = (x + y + 3z)
2
2yz 8z
2
= (x + y + 3z)
2
8
z +
y
8
2
+
1
8
y
2
.
We now see
φ
x
y
z
,
x
0
y
0
z
0
= (x + y + 3z)(x
0
+ y
0
+ 3z
0
) 8
z +
y
8
z
0
+
y
0
8
+
1
8
yy
0
.
Why do we know this? This is clearly a symmetric bilinear form, and this also
clearly induces the
q
given above. By uniqueness, we know that this is the right
symmetric bilinear form.
We now use this form to find our f
1
, f
2
, f
3
such that φ(f
i
, f
j
) = δ
ij
.
To do so, we just solve the equations
x + y + 3z = 1
z +
1
8
y = 0
y = 0.
This gives our first vector as
f
1
=
1
0
0
.
We then solve
x + y + 3z = 0
z +
1
8
y = 1
y = 0.
So we have
f
2
=
3
0
1
.
Finally, we solve
x + y + 3z = 0
z +
1
8
y = 0
y = 1.
This gives
f
3
=
5/8
1
1/8.
.
Then we can see that the result follows, and we get
q(af
1
+ bf
2
+ cf
3
) = a
2
8b
2
+
1
8
c
2
.
We see that the diagonal matrix we get is not unique. We can re-scale our
basis by any constant, and get an equivalent expression.
Theorem. Let φ be a symmetric bilinear form over a complex vector space V .
Then there exists a basis (v
1
, ··· , v
m
) for V such that φ is represented by
I
r
0
0 0
with respect to this basis, where r = r(φ).
Proof.
We’ve already shown that there exists a basis (
e
1
, ··· , e
n
) such that
φ
(
e
i
, e
j
) =
λ
i
δ
ij
for some
λ
ij
. By reordering the
e
i
, we can assume that
λ
1
, ··· , λ
r
6= 0 and λ
r+1
, ··· , λ
n
= 0.
For each 1
i r
, there exists some
µ
i
such that
µ
2
i
=
λ
i
. For
r
+ 1
r n
,
we let µ
i
= 1 (or anything non-zero). We define
v
i
=
e
i
µ
i
.
Then
φ(v
i
, v
j
) =
1
µ
i
µ
j
φ(e
i
, e
j
) =
(
0 i 6= j or i = j > r
1 i = j < r.
So done.
Note that it follows that for the corresponding quadratic form q, we have
q
n
X
i=1
a
i
v
i
!
=
r
X
i=1
a
2
i
.
Corollary.
Every symmetric
A Mat
n
(
C
) is congruent to a unique matrix of
the form
I
r
0
0 0
.
Now this theorem is a bit too strong, and we are going to fix that next lecture,
by talking about Hermitian forms and sesquilinear forms. Before that, we do
the equivalent result for real vector spaces.
Theorem.
Let
φ
be a symmetric bilinear form of a finite-dimensional vector
space over
R
. Then there exists a basis (
v
1
, ··· , v
n
) for
V
such that
φ
is
represented
I
p
I
q
0
,
with
p
+
q
=
r
(
φ
),
p, q
0. Equivalently, the corresponding quadratic forms is
given by
q
n
X
i=1
a
i
v
i
!
=
p
X
i=1
a
2
i
p+q
X
j=p+1
a
2
j
.
Note that we have seen these things in special relativity, where the Minkowski
inner product is given by the symmetric bilinear form represented by
1 0 0 0
0 1 0 0
0 0 1 0
0 0 0 1
,
in units where c = 1.
Proof.
We’ve already shown that there exists a basis (
e
1
, ··· , e
n
) such that
φ(e
i
, e
j
) = λ
i
δ
ij
for some λ
1
, ··· , λ
n
R. By reordering, we may assume
λ
i
> 0 1 i p
λ
i
< 0 p + 1 i r
λ
i
= 0 i > r
We let µ
i
be defined by
µ
i
=
λ
i
1 i p
λ
i
p + 1 i r
1 i > r
Defining
v
i
=
1
µ
i
e
i
,
we find that φ is indeed represented by
I
p
I
q
0
,
We will later show that this form is indeed unique. Before that, we will have
a few definitions, that really only make sense over R.
Definition
(Positive/negative (semi-)definite)
.
Let
φ
be a symmetric bilinear
form on a finite-dimensional real vector space V . We say
(i) φ is positive definite if φ(v, v) > 0 for all v V \{0}.
(ii) φ is positive semi-definite if φ(v, v) 0 for all v V .
(iii) φ is negative definite if φ(v, v) < 0 for all v V \{0}.
(iv) φ is negative semi-definite if φ(v, v) 0 for all v V .
We are going to use these notions to prove uniqueness. It is easy to see that
if
p
= 0 and
q
=
n
, then we are negative definite; if
p
= 0 and
q 6
=
n
, then we
are negative semi-definite etc.
Example. Let φ be a symmetric bilinear form on R
n
represented by
I
p
0
0 0
np
.
Then φ is positive semi-definite. φ is positive definite if and only if n = p.
If instead φ is represented by
I
p
0
0 0
np
,
then φ is negative semi-definite. φ is negative definite precisely if n = q.
We are going to use this to prove the uniqueness part of our previous theorem.
Theorem
(Sylvester’s law of inertia)
.
Let
φ
be a symmetric bilinear form on a
finite-dimensional real vector space
V
. Then there exists unique non-negative
integers p, q such that φ is represented by
I
p
0 0
0 I
q
0
0 0 0
with respect to some basis.
Proof.
We have already proved the existence part, and we just have to prove
uniqueness. To do so, we characterize
p
and
q
in a basis-independent way. We
already know that
p
+
q
=
r
(
φ
) does not depend on the basis. So it suffices to
show p is unique.
To see that
p
is unique, we show that
p
is the largest dimension of a subspace
P V such that φ|
P ×P
is positive definite.
First we show we can find such at P . Suppose φ is represented by
I
p
0 0
0 I
q
0
0 0 0
with respect to (
e
1
, ··· , e
n
). Then
φ
restricted to
he
1
, ··· , e
p
i
is represented by
I
p
with respect to e
1
, ··· , e
p
. So φ restricted to this is positive definite.
Now suppose
P
is any subspace of
V
such that
φ|
P ×P
is positive definite. To
show
P
has dimension at most
p
, we find a subspace complementary to
P
with
dimension n p.
Let Q = he
p+1
, ··· , e
n
i. Then φ restricted to Q × Q is represented by
I
q
0
0 0
.
Now if
v P Q \{
0
}
, then
φ
(
v, v
)
>
0 since
v P \{
0
}
and
φ
(
v, v
)
0 since
v Q, which is a contradiction. So P Q = 0.
We have
dim V dim(P + Q) = dim P + dim Q = dim P + (n p).
Rearranging gives
dim P p.
A similar argument shows that
q
is the maximal dimension of a subspace
Q V
such that φ|
Q×Q
is negative definite.
Definition
(Signature)
.
The signature of a bilinear form
φ
is the number
p q
,
where p and q are as above.
Of course, we can recover p and q from the signature and the rank of φ.
Corollary.
Every real symmetric matrix is congruent to precisely one matrix
of the form
I
p
0 0
0 I
q
0
0 0 0
.
7.2 Hermitian form
The above result was nice for real vector spaces. However, if
φ
is a bilinear form
on a
C
-vector space
V
, then
φ
(
iv, iv
) =
φ
(
v, v
). So there can be no good
notion of positive definiteness for complex bilinear forms. To make them work
for complex vector spaces, we need to modify the definition slightly to obtain
Hermitian forms.
Definition
(Sesquilinear form)
.
Let
V, W
be complex vector spaces. Then a
sesquilinear form is a function φ : V ×W C such that
(i) φ(λv
1
+ µv
2
, w) =
¯
λφ(v
1
, w) + ¯µφ(v
2
, w).
(ii) φ(v, λw
1
+ µw
2
) = λφ(v, w
1
) + µφ(vw
2
).
for all v, v
1
, v
2
V , w, w
1
, w
2
W and λ, µ C.
Note that some people have an opposite definition, where we have linearity
in the first argument and conjugate linearity in the second.
These are called sesquilinear since “sesqui” means “one and a half”, and this
is linear in the second argument and “half linear” in the first.
Alternatively, to define a sesquilinear form, we can define a new complex
vector space
¯
V structure on V by taking the same abelian group (i.e. the same
underlying set and addition), but with the scalar multiplication
C ×
¯
V
¯
V
defined as
(λ, v) 7→
¯
λv.
Then a sesquilinear form on
V × W
is a bilinear form on
¯
V × W
. Alternatively,
this is a linear map W
¯
V
.
Definition
(Representation of sesquilinear form)
.
Let
V, W
be finite-dimensional
complex vector spaces with basis (
v
1
, ··· , v
n
) and (
w
1
, ··· , w
m
) respectively,
and
φ
:
V × W C
be a sesquilinear form. Then the matrix representing
φ
with respect to these bases is
A
ij
= φ(v
i
, w
j
).
for 1 i n, 1 j m.
As usual, this determines the whole sesquilinear form. This follows from
the analogous fact for the bilinear form on
¯
V ×W C
. Let
v
=
P
λ
i
v
i
and
W =
P
µ
j
w
j
. Then we have
φ(v, w) =
X
i,j
λ
i
µ
j
φ(v
i
, w
j
) = λ
Aµ.
We now want the right definition of symmetric sesquilinear form. We cannot
just require
φ
(
v, w
) =
φ
(
w, v
), since
φ
is linear in the second variable and
conjugate linear on the first variable. So in particular, if
φ
(
v, w
)
6
= 0, we have
φ(iv, w) 6= φ(v, iw).
Definition
(Hermitian sesquilinear form)
.
A sesquilinear form on
V × V
is
Hermitian if
φ(v, w) = φ(w, v).
Note that if
φ
is Hermitian, then
φ
(
v, v
) =
φ(v, v) R
for any
v V
. So
it makes sense to ask if it is positive or negative. Moreover, for any complex
number λ, we have
φ(λv, λv) = |λ|
2
φ(v, v).
So multiplying by a scalar does not change the sign. So it makes sense to talk
about positive (semi-)definite and negative (semi-)definite Hermitian forms.
We will prove results analogous to what we had for real symmetric bilinear
forms.
Lemma.
Let
φ
:
V × V C
be a sesquilinear form on a finite-dimensional
vector space over
C
, and (
e
1
, ··· , e
n
) a basis for
V
. Then
φ
is Hermitian if and
only if the matrix A representing φ is Hermitian (i.e. A = A
).
Proof. If φ is Hermitian, then
A
ij
= φ(e
i
, e
j
) = φ(e
j
, e
i
) = A
ij
.
If A is Hermitian, then
φ
X
λ
i
e
i
,
X
µ
j
e
j
= λ
= µ
A
λ = φ
X
µ
j
e
j
,
X
λ
j
e
j
.
So done.
Proposition
(Change of basis)
.
Let
φ
be a Hermitian form on a finite di-
mensional vector space
V
; (
e
1
, ··· , e
n
) and (
v
1
, ··· , v
n
) are bases for
V
such
that
v
i
=
n
X
k=1
P
ki
e
k
;
and
A, B
represent
φ
with respect to (
e
1
, . . . , e
n
) and (
v
1
, ··· , v
n
) respectively.
Then
B = P
AP.
Proof. We have
B
ij
= φ(v
i
, v
j
)
= φ
X
P
ki
e
k
,
X
P
`j
e
`
=
n
X
k,`=1
¯
P
ki
P
`j
A
k`
= (P
AP )
ij
.
Lemma
(Polarization identity (again))
.
A Hermitian form
φ
on
V
is determined
by the function ψ : v 7→ φ(v, v).
The proof this time is slightly more involved.
Proof. We have the following:
ψ(x + y) = φ(x, x) + φ(x, y) + φ(y, x) + φ(y, y)
ψ(x y) = φ(x, x) + φ(x, y) + φ(y, x) φ(y, y)
(x iy) = (x, x) + φ(x, y) φ(y, x) + (y, y)
(x + iy) = (x, x) + φ(x, y) φ(y, x) (y, y)
So
φ(x, y) =
1
4
(ψ(x + y) ψ(x y) + (x iy) (x + iy)).
Theorem
(Hermitian form of Sylvester’s law of inertia)
.
Let
V
be a finite-
dimensional complex vector space and
φ
a hermitian form on
V
. Then there
exists unique non-negative integers p and q such that φ is represented by
I
p
0 0
0 I
q
0
0 0 0
with respect to some basis.
Proof. Same as for symmetric forms over R.
8 Inner product spaces
Welcome to the last chapter, where we discuss inner products. Technically, an
inner product is just a special case of a positive-definite symmetric bilinear or
hermitian form. However, they are usually much more interesting and useful.
Many familiar notions such as orthogonality only make sense when we have an
inner product.
In this chapter, we will adopt the convention that
F
always means either
R
or C, since working with other fields doesn’t make much sense here.
8.1 Definitions and basic properties
Definition
(Inner product space)
.
Let
V
be a vector space. An inner product
on
V
is a positive-definite symmetric bilinear/hermitian form. We usually write
(x, y) instead of φ(x, y).
A vector space equipped with an inner product is an inner product space.
We will see that if we have an inner product, then we can define lengths and
distances in a sensible way.
Example.
(i) R
n
or C
n
with the usual inner product
(x, y) =
n
X
i=1
¯x
i
y
i
forms an inner product space.
In some sense, this is the only inner product on finite-dimensional spaces,
by Sylvester’s law of inertia. However, we would not like to think so, and
instead work with general inner products.
(ii)
Let
C
([0
,
1]
, F
) be the vector space of real/complex valued continuous
functions on [0, 1]. Then the following is an inner product:
(f, g) =
Z
1
0
¯
f(t)g(t) dt.
(iii)
More generally, for any
w
: [0
,
1]
R
+
continuous, we can define the inner
product on C([0, 1], F) as
(f, g) =
Z
1
0
w(t)
¯
f(t)g(t) dt.
If V is an inner product space, we can define a norm on V by
kvk =
p
(v, v).
This is just the usual notion of norm on
R
n
and
C
n
. This gives the notion of
length in inner product spaces. Note that
kvk >
0 with equality if and only if
v = 0.
Note also that the norm
k · k
determines the inner product by the polarization
identity.
We want to see that this indeed satisfies the definition of a norm, as you might
have seen from Analysis II. To prove this, we need to prove the Cauchy-Schwarz
inequality.
Theorem
(Cauchy-Schwarz inequality)
.
Let
V
be an inner product space and
v, w V . Then
|(v, w)| kvkkwk.
Proof.
If
w
= 0, then this is trivial. Otherwise, since the norm is positive
definite, for any λ, we get
0 (v λw, v λw) = (v, v)
¯
λ(w, v) λ(v, w) + |λ|
2
(w, w).
We now pick a clever value of λ. We let
λ =
(w, v)
(w, w)
.
Then we get
0 (v, v)
|(w, v)|
2
(w, w)
|(w, v)|
2
(w, w)
+
|(w, v)|
2
(w, w)
.
So we get
|(w, v)|
2
(v, v)(w, w).
So done.
With this, we can prove the triangle inequality.
Corollary
(Triangle inequality)
.
Let
V
be an inner product space and
v, w V
.
Then
kv + wk kv k + kwk.
Proof. We compute
kv + wk
2
= (v + w, v + w)
= (v, v) + (v, w) + (w, v) + (w, w)
kvk
2
+ 2kvkkwk + kwk
2
= (kvk + kwk)
2
.
So done.
The next thing we do is to define orthogonality. This generalizes the notion
of being “perpendicular”.
Definition
(Orthogonal vectors)
.
Let
V
be an inner product space. Then
v, w V are orthogonal if (v, w) = 0.
Definition
(Orthonormal set)
.
Let
V
be an inner product space. A set
{v
i
:
i I} is an orthonormal set if for any i, j I, we have
(v
i
, v
j
) = δ
ij
It should be clear that an orthonormal set must be linearly independent.
Definition
(Orthonormal basis)
.
Let
V
be an inner product space. A subset of
V is an orthonormal basis if it is an orthonormal set and is a basis.
In an inner product space, we almost always want orthonormal basis only. If
we pick a basis, we should pick an orthonormal one.
However, we do not know there is always an orthonormal basis, even in the
finite-dimensional case. Also, given an orthonormal set, we would like to extend
it to an orthonormal basis. This is what we will do later.
Before that, we first note that given an orthonormal basis, it is easy to find
the coordinates of any vector in this basis. Suppose
V
is a finite-dimensional
inner product space with an orthonormal basis v
1
, ··· , v
n
. Given
v =
n
X
i=1
λ
i
v
i
,
we have
(v
j
, v) =
n
X
i=1
λ
i
(v
j
, v
i
) = λ
j
.
So v V can always be written as
n
X
i=1
(v
i
, v)v
i
.
Lemma
(Parseval’s identity)
.
Let
V
be a finite-dimensional inner product space
with an orthonormal basis v
1
, ··· , v
n
, and v, w V . Then
(v, w) =
n
X
i=1
(v
i
, v)(v
i
, w).
In particular,
kvk
2
=
n
X
i=1
|(v
i
, v)|
2
.
This is something we’ve seen in IB Methods, for infinite dimensional spaces.
However, we will only care about finite-dimensional ones now.
Proof.
(v, w) =
n
X
i=1
(v
i
, v)v
i
,
n
X
j=1
(v
j
, w)v
j
=
n
X
i,j=1
(v
i
, v)(v
j
, w)(v
i
, v
j
)
=
n
X
i,j=1
(v
i
, v)(v
j
, w)δ
ij
=
n
X
i=1
(v
i
, v)(v
i
, w).
8.2 Gram-Schmidt orthogonalization
As mentioned, we want to make sure every vector space has an orthonormal
basis, and we can extend any orthonormal set to an orthonormal basis, at least
in the case of finite-dimensional vector spaces. The idea is to start with an
arbitrary basis, which we know exists, and produce an orthonormal basis out of
it. The way to do this is the Gram-Schmidt process.
Theorem
(Gram-Schmidt process)
.
Let
V
be an inner product space and
e
1
, e
2
, ···
a linearly independent set. Then we can construct an orthonormal set
v
1
, v
2
, ··· with the property that
hv
1
, ··· , v
k
i = he
1
, ··· , e
k
i
for every k.
Note that we are not requiring the set to be finite. We are just requiring it
to be countable.
Proof.
We construct it iteratively, and prove this by induction on
k
. The base
case k = 0 is contentless.
Suppose we have already found v
1
, ··· , v
k
that satisfies the properties. We
define
u
k+1
= e
k+1
k
X
i=1
(v
i
, e
i+1
)v
i
.
We want to prove that this is orthogonal to all the other
v
i
’s for
i k
. We have
(v
j
, u
k+1
) = (v
j
, e
k+1
)
k
X
i=1
(v
i
, e
k+1
)δ
ij
= (v
j
, e
k+1
) (v
j
, e
k+1
) = 0.
So it is orthogonal.
We want to argue that u
k+1
is non-zero. Note that
hv
1
, ··· , v
k
, u
k+1
i = hv
1
, ··· , v
k
, e
k+1
i
since we can recover
e
k+1
from
v
1
, ··· , v
k
and
u
k+1
by construction. We also
know
hv
1
, ··· , v
k
, e
k+1
i = he
1
, ··· , e
k
, e
k+1
i
by assumption. We know
he
1
, ··· , e
k
, e
k+1
i
has dimension
k
+ 1 since the
e
i
are
linearly independent. So we must have
u
k+1
non-zero, or else
hv
1
, ··· , v
k
i
will
be a set of size
k
spanning a space of dimension
k
+ 1, which is clearly nonsense.
Therefore, we can define
v
k+1
=
u
k+1
ku
k+1
k
.
Then
v
1
, ··· , v
k+1
is orthonormal and
hv
1
, ··· , v
k+1
i
=
he
1
, ··· , e
k+1
i
as re-
quired.
Corollary.
If
V
is a finite-dimensional inner product space, then any orthonor-
mal set can be extended to an orthonormal basis.
Proof.
Let
v
1
, ··· , v
k
be an orthonormal set. Since this is linearly independent,
we can extend it to a basis (v
1
, ··· , v
k
, x
k+1
, ··· , x
n
).
We now apply the Gram-Schmidt process to this basis to get an orthonormal
basis of
V
, say (
u
1
, ··· , u
n
). Moreover, we can check that the process does not
modify our v
1
, ··· , v
k
, i.e. u
i
= v
i
for 1 i k. So done.
Definition
(Orthogonal internal direct sum)
.
Let
V
be an inner product space
and
V
1
, V
2
V
. Then
V
is the orthogonal internal direct sum of
V
1
and
V
2
if it
is a direct sum and V
1
and V
2
are orthogonal. More precisely, we require
(i) V = V
1
+ V
2
(ii) V
1
V
2
= 0
(iii) (v
1
, v
2
) = 0 for all v
1
V
1
and v
2
V
2
.
Note that condition (iii) implies (ii), but we write it for the sake of explicitness.
We write V = V
1
V
2
.
Definition
(Orthogonal complement)
.
If
W V
is a subspace of an inner
product space V , then the orthogonal complement of W in V is the subspace
W
= {v V : (v, w) = 0, w W }.
It is true that the orthogonal complement is a complement and orthogonal,
i.e. V is the orthogonal direct sum of W and W
.
Proposition.
Let
V
be a finite-dimensional inner product space, and
W V
.
Then
V = W W
.
Proof.
There are three things to prove, and we know (iii) implies (ii). Also, (iii)
is obvious by definition of W
. So it remains to prove (i), i.e. V = W + W
.
Let w
1
, ··· , w
k
be an orthonormal basis for W , and pick v V . Now let
w =
k
X
i=1
(w
i
, v)w
i
.
Clearly, we have
w W
. So we need to show
v w W
. For each
j
, we can
compute
(w
j
, v w) = (w
j
, v)
k
X
i=1
(w
i
, v)(w
j
, w
i
)
= (w
j
, v)
k
X
i=1
(w
i
, v)δ
ij
= 0.
Hence for any λ
i
, we have
X
λ
j
w
j
, v w
= 0.
So we have v w W
. So done.
Definition
(Orthogonal external direct sum)
.
Let
V
1
, V
2
be inner product spaces.
The orthogonal external direct sum of
V
1
and
V
2
is the vector space
V
1
V
2
with
the inner product defined by
(v
1
+ v
2
, w
1
+ w
2
) = (v
1
, w
1
) + (v
2
, w
2
),
with v
1
, w
1
V
1
, v
2
, w
2
V
2
.
Here we write v
1
+ v
2
V
1
V
2
instead of (v
1
, v
2
) to avoid confusion.
This external direct sum is equivalent to the internal direct sum of
{
(
v
1
, 0
) :
v
1
V
1
} and {(0, v
2
) : v
2
V
2
}.
Proposition.
Let
V
be a finite-dimensional inner product space and
W V
.
Let (
e
1
, ··· , e
k
) be an orthonormal basis of
W
. Let
π
be the orthonormal
projection of
V
onto
W
, i.e.
π
:
V W
is a function that satisfies
ker π
=
W
,
π|
W
= id. Then
(i) π is given by the formula
π(v) =
k
X
i=1
(e
i
, v)e
i
.
(ii) For all v V, w W , we have
kv π(v)k kv wk,
with equality if and only if
π
(
v
) =
w
. This says
π
(
v
) is the point on
W
that is closest to v.
W
w
v
π(v)
Proof.
(i) Let v V , and define
w =
X
i=1
(e
i
, v)e
i
.
We want to show this is
π
(
v
). We need to show
v w W
. We can
compute
(e
j
, v w) = (e
j
, v)
k
X
i=1
(e
i
, v)(e
j
, e
i
) = 0.
So v w is orthogonal to every basis vector in w, i.e. v w W
.So
π(v) = π(w) + π(v w) = w
as required.
(ii)
This is just Pythagoras’ theorem. Note that if
x
and
y
are orthogonal,
then
kx + yk
2
= (x + y, x + y)
= (x, x) + (x, y) + (y, x) + (y.y)
= kxk
2
+ kyk
2
.
We apply this to our projection. For any w W , we have
kv wk
2
= kv π(v)k
2
+ kπ(v) wk
2
kv π(v)k
2
with equality if and only if kπ(v) wk = 0, i.e. π(v) = w.
8.3 Adjoints, orthogonal and unitary maps
Adjoints
Lemma.
Let
V
and
W
be finite-dimensional inner product spaces and
α
:
V
W
is a linear map. Then there exists a unique linear map
α
:
W V
such that
(αv, w) = (v, α
w) ()
for all v V , w W .
Proof.
There are two parts. We have to prove existence and uniqueness. We’ll
first prove it concretely using matrices, and then provide a conceptual reason of
what this means.
Let (
v
1
, ··· , v
n
) and (
w
1
, ··· , w
m
) be orthonormal basis for
V
and
W
.
Suppose α is represented by A.
To show uniqueness, suppose
α
:
W V
satisfies (
αv, w
) = (
v, α
w
) for
all v V , w W , then for all i, j, by definition, we know
(v
i
, α
(w
j
)) = (α(v
i
), w
j
)
=
X
k
A
ki
w
k
, w
j
!
=
X
k
¯
A
ki
(w
k
, w
j
) =
¯
A
ji
.
So we get
α
(w
j
) =
X
i
(v
i
, α
(w
j
))v
i
=
X
i
¯
A
ji
v
i
.
Hence α
must be represented by A
. So α
is unique.
To show existence, all we have to do is to show
A
indeed works. Now let
α
be represented by
A
. We can compute the two sides of (
) for arbitrary
v, w
.
We have
α
X
λ
i
v
i
,
X
µ
j
w
j
=
X
i,j
¯
λ
i
µ
j
(α(v
i
), w
j
)
=
X
i,j
¯
λ
i
µ
j
X
k
A
ki
w
k
, w
j
!
=
X
i,j
¯
λ
i
¯
A
ji
µ
j
.
We can compute the other side and get
X
λ
i
v
i
, α
X
µ
j
w
j

=
X
i,j
¯
λ
i
µ
j
v
i
,
X
k
A
kj
v
k
!
=
X
i,j
¯
λ
i
¯
A
ji
µ
j
.
So done.
What does this mean, conceptually? Note that the inner product
V
defines
an isomorphism
V
¯
V
by
v 7→
(
·, v
). Similarly, we have an isomorphism
W
¯
W
. We can then put them in the following diagram:
V W
¯
V
¯
W
α
=
=
α
Then
α
is what fills in the dashed arrow. So
α
is in some sense the “dual” of
the map α.
Definition (Adjoint). We call the map α
the adjoint of α.
We have just seen that if
α
is represented by
A
with respect to some or-
thonormal bases, then α
is represented by A
.
Definition
(Self-adjoint)
.
Let
V
be an inner product space, and
α End
(
V
).
Then α is self-adjoint if α = α
, i.e.
(α(v), w) = (v, α(w))
for all v, w.
Thus if
V
=
R
n
with the usual inner product, then
A Mat
n
(
R
) is self-
adjoint if and only if it is symmetric, i.e.
A
=
A
T
. If
V
=
C
n
with the usual
inner product, then
A Mat
n
(
C
) is self-adjoint if and only if
A
is Hermitian,
i.e. A = A
.
Self-adjoint endomorphisms are important, as you may have noticed from IB
Quantum Mechanics. We will later see that these have real eigenvalues with an
orthonormal basis of eigenvectors.
Orthogonal maps
Another important class of endomorphisms is those that preserve lengths. We
will first do this for real vector spaces, since the real and complex versions have
different names.
Definition
(Orthogonal endomorphism)
.
Let
V
be a real inner product space.
Then α End(V ) is orthogonal if
(α(v), α(w)) = (v, w)
for all v, w V .
By the polarization identity,
α
is orthogonal if and only if
kα
(
v
)
k
=
kvk
for
all v V .
A real square matrix (as an endomorphism of
R
n
with the usual inner product)
is orthogonal if and only if its columns are an orthonormal set.
There is also an alternative way of characterizing these orthogonal maps.
Lemma.
Let
V
be a finite-dimensional space and
α End
(
V
). Then
α
is
orthogonal if and only if α
1
= α
.
Proof. () Suppose α
1
= α
. If α
1
= α
, then
(αv, αv) = (v, α
αv) = (v, α
1
αv) = (v, v).
(
) If
α
is orthogonal and (
v
1
, ··· , v
n
) is an orthonormal basis for
V
, then for
1 i, j n, we have
δ
ij
= (v
i
, v
j
) = (αv
i
, αv
j
) = (v
i
, α
αv
j
).
So we know
α
α(v
j
) =
n
X
i=1
(v
i
, α
αv
j
))v
i
= v
j
.
So by linearity of α
α, we know α
α = id
V
. So α
= α
1
.
Corollary. α End
(
V
) is orthogonal if and only if
α
is represented by an
orthogonal matrix, i.e. a matrix
A
such that
A
T
A
=
AA
T
=
I
, with respect to
any orthonormal basis.
Proof.
Let (
e
1
, ··· , e
n
) be an orthonormal basis for
V
. Then suppose
α
is
represented by
A
. So
α
is represented by
A
T
. Then
A
=
A
1
if and only if
AA
T
= A
T
A = I.
Definition
(Orthogonal group)
.
Let
V
be a real inner product space. Then the
orthogonal group of V is
O(V ) = {α End(V ) : α is orthogonal}.
It follows from the fact that
α
=
α
1
that
α
is invertible, and it is clear
from definition that O(
V
) is closed under multiplication and inverses. So this is
indeed a group.
Proposition.
Let
V
be a finite-dimensional real inner product space and
(e
1
, ··· , e
n
) is an orthonormal basis of V . Then there is a bijection
O(V ) {orthonormal basis for V }
α 7→ (α(e
1
, ··· , e
n
)).
This is analogous to our result for general vector spaces and general bases,
where we replace O(V ) with GL(V ).
Proof. Same as the case for general vector spaces and general bases.
Unitary maps
We are going to study the complex version of orthogonal maps, known as unitary
maps. The proofs are almost always identical to the real case, and we will not
write the proofs again.
Definition
(Unitary map)
.
Let
V
be a finite-dimensional complex vector space.
Then α End(V ) is unitary if
(α(v), α(w)) = (v, w)
for all v, w V .
By the polarization identity,
α
is unitary if and only if
kα
(
v
)
k
=
kvk
for all
v V .
Lemma.
Let
V
be a finite dimensional complex inner product space and
α
End(V ). Then α is unitary if and only if α is invertible and α
= α
1
.
Corollary. α End
(
V
) is unitary if and only if
α
is represented by a unitary
matrix A with respect to any orthonormal basis, i.e. A
1
= A
.
Definition
(Unitary group)
.
Let
V
be a finite-dimensional complex inner prod-
uct space. Then the unitary group of V is
U(V ) = {α End(V ) : α is unitary}.
Proposition.
Let
V
be a finite-dimensional complex inner product space. Then
there is a bijection
U(V ) {orthonormal basis of V }
α 7→ {α(e
1
), ··· , α(e
n
)}.
8.4 Spectral theory
We are going to classify matrices in inner product spaces. Recall that for general
vector spaces, what we effectively did was to find the orbits of the conjugation
action of
GL
(
V
) on
Mat
n
(
F
). If we have inner product spaces, we will want to
look at the action of
O
(
V
) or
U
(
V
) on
Mat
n
(
F
). In a more human language,
instead of allowing arbitrary basis transformations, we only allow transforming
between orthonormal basis.
We are not going to classify all endomorphisms, but just self-adjoint and
orthogonal/unitary ones.
Lemma.
Let
V
be a finite-dimensional inner product space, and
α End
(
V
)
self-adjoint. Then
(i) α has a real eigenvalue, and all eigenvalues of α are real.
(ii) Eigenvectors of α with distinct eigenvalues are orthogonal.
Proof. We are going to do real and complex cases separately.
(i)
Suppose first
V
is a complex inner product space. Then by the fundamental
theorem of algebra,
α
has an eigenvalue, say
λ
. We pick
v V \ {
0
}
such
that αv = λv. Then
¯
λ(v, v) = (λv, v) = (αv, v) = (v, αv) = (v, λv) = λ(v, v).
Since v 6= 0, we know (v, v) 6= 0. So λ =
¯
λ.
For the real case, we pretend we are in the complex case. Let
e
1
, ··· , e
n
be
an orthonormal basis for
V
. Then
α
is represented by a symmetric matrix
A
(with respect to this basis). Since real symmetric matrices are Hermitian
viewed as complex matrices, this gives a self-adjoint endomorphism of
C
n
.
By the complex case,
A
has real eigenvalues only. But the eigenvalues of
A are the eigenvalues of α and M
A
(t) = M
α
(t). So done.
Alternatively, we can prove this without reducing to the complex case. We
know every irreducible factor of
M
α
(
t
) in
R
[
t
] must have degree 1 or 2,
since the roots are either real or come in complex conjugate pairs. Suppose
f(t) were an irreducible factor of degree 2. Then
m
α
f
(α) 6= 0
since it has degree less than the minimal polynomial. So there is some
v V such that
M
α
f
(α)(v) 6= 0.
So it must be that
f
(
α
)(
v
) =
0
. Let
U
=
hv, α
(
v
)
i
. Then this is an
α-invariant subspace of V since f has degree 2.
Now
α|
U
End
(
U
) is self-adjoint. So if (
e
1
, e
2
) is an orthonormal basis of
U, then α is represented by a real symmetric matrix, say
a b
b a
But then
χ
α|
U
(
t
) = (
t a
)
2
b
2
, which has real roots, namely
a ± b
. This
is a contradiction, since M
α|
U
= f, but f is irreducible.
(ii)
Now suppose
αv
=
λv
,
αw
=
µw
and
λ 6
=
µ
. We need to show (
v, w
) = 0.
We know
(αv, w) = (v, αw)
by definition. This then gives
λ(v, w) = µ(v, w)
Since λ 6= µ, we must have (v, w) = 0.
Theorem.
Let
V
be a finite-dimensional inner product space, and
α End
(
V
)
self-adjoint. Then V has an orthonormal basis of eigenvectors of α.
Proof.
By the previous lemma,
α
has a real eigenvalue, say
λ
. Then we can find
an eigenvector v V \{0} such that αv = λv.
Let U = hvi
. Then we can write
V = hvi U.
We now want to prove α sends U into U. Suppose u U . Then
(v, α(u)) = (αv, u) = λ(v, u) = 0.
So α(u) hvi
= U. So α|
U
End(U ) and is self-adjoint.
By induction on
dim V
,
U
has an orthonormal basis (
v
2
, ··· , v
n
) of
α
eigen-
vectors. Now let
v
1
=
v
kvk
.
Then (v
1
, v
2
, ··· , v
n
) is an orthonormal basis of eigenvectors for α.
Corollary.
Let
V
be a finite-dimensional vector space and
α
self-adjoint. Then
V is the orthogonal (internal) direct sum of its α-eigenspaces.
Corollary.
Let
A Mat
n
(
R
) be symmetric. Then there exists an orthogonal
matrix P such that P
T
AP = P
1
AP is diagonal.
Proof.
Let (
·, ·
) be the standard inner product on
R
n
. Then
A
is self-adjoint
as an endomorphism of
R
n
. So
R
n
has an orthonormal basis of eigenvectors for
A, say (v
1
, ··· , v
n
). Taking P = (v
1
v
2
··· v
n
) gives the result.
Corollary.
Let
V
be a finite-dimensional real inner product space and
ψ
:
V × V R
a symmetric bilinear form. Then there exists an orthonormal basis
(
v
1
, ··· , v
n
) for
V
with respect to which
ψ
is represented by a diagonal matrix.
Proof.
Let (
u
1
, ··· , u
n
) be any orthonormal basis for
V
. Then
ψ
is represented
by a symmetric matrix
A
. Then there exists an orthogonal matrix
P
such that
P
T
AP
is diagonal. Now let
v
i
=
P
P
ki
u
k
. Then (
v
1
, ··· , v
n
) is an orthonormal
basis since
(v
i
, v
j
) =
X
P
ki
u
k
,
X
P
`j
u
`
=
X
P
T
ik
(u
k
, u
`
)P
`j
= [P
T
P ]
ij
= δ
ij
.
Also, ψ is represented by P
T
AP with respect to (v
1
, ··· , v
n
).
Note that the diagonal values of
P
T
AP
are just the eigenvalues of
A
. So the
signature of
ψ
is just the number of positive eigenvalues of
A
minus the number
of negative eigenvalues of A.
Corollary.
Let
V
be a finite-dimensional real vector space and
φ, ψ
symmetric
bilinear forms on
V
such that
φ
is positive-definite. Then we can find a basis
(
v
1
, ··· , v
n
) for
V
such that both
φ
and
ψ
are represented by diagonal matrices
with respect to this basis.
Proof.
We use
φ
to define an inner product. Choose an orthonormal basis for
V
(equipped with
φ
) (
v
1
, ··· , v
n
) with respect to which
ψ
is diagonal. Then
φ
is
represented by I with respect to this basis, since ψ(v
i
, v
j
) = δ
ij
. So done.
Corollary.
If
A, B Mat
n
(
R
) are symmetric and
A
is positive definitive (i.e.
v
T
Av >
0 for all
v R
n
\ {
0
}
). Then there exists an invertible matrix
Q
such
that Q
T
AQ and Q
T
BQ are both diagonal.
We can deduce similar results for complex finite-dimensional vector spaces,
with the same proofs. In particular,
Proposition.
(i)
If
A Mat
n
(
C
) is Hermitian, then there exists a unitary matrix
U
Mat
n
(C) such that
U
1
AU = U
AU
is diagonal.
(ii)
If
ψ
is a Hermitian form on a finite-dimensional complex inner product
space V , then there is an orthonormal basis for V diagonalizing ψ.
(iii)
If
φ, ψ
are Hermitian forms on a finite-dimensional complex vector space
and
φ
is positive definite, then there exists a basis for which
φ
and
ψ
are
diagonalized.
(iv)
Let
A, B Mat
n
(
C
) be Hermitian, and
A
positive definitive (i.e.
v
Av >
0
for
v V \{
0
}
). Then there exists some invertible
Q
such that
Q
AQ
and
Q
BQ are diagonal.
That’s all for self-adjoint matrices. How about unitary matrices?
Theorem.
Let
V
be a finite-dimensional complex vector space and
α U
(
V
)
be unitary. Then V has an orthonormal basis of α eigenvectors.
Proof.
By the fundamental theorem of algebra, there exists
v V \ {
0
}
and
λ C such that αv = λv. Now consider W = hvi
. Then
V = W hvi.
We want to show
α
restricts to a (unitary) endomorphism of
W
. Let
w W
.
We need to show α(w) is orthogonal to v. We have
(αw, v) = (w, α
1
v) = (w, λ
1
v) = 0.
So
α
(
w
)
W
and
α|
W
End
(
W
). Also,
α|
W
is unitary since
α
is. So by
induction on
dim V
,
W
has an orthonormal basis of
α
eigenvectors. If we add
v/kvk
to this basis, we get an orthonormal basis of
V
itself comprised of
α
eigenvectors.
This theorem and the analogous one for self-adjoint endomorphisms have a
common generalization, at least for complex inner product spaces. The key fact
that leads to the existence of an orthonormal basis of eigenvectors is that
α
and
α
commute. This is clearly a necessary condition, since if
α
is diagonalizable, then
α
is diagonal in the same basis (since it is just the transpose (and conjugate)),
and hence they commute. It turns out this is also a sufficient condition, as you
will show in example sheet 4.
However, we cannot generalize this in the real orthogonal case. For example,
cos θ sin θ
sin θ cos θ
O(R
2
)
cannot be diagonalized (if
θ 6∈ πZ
). However, in example sheet 4, you will find a
classification of
O
(
V
), and you will see that the above counterexample is the
worst that can happen in some sense.