0% found this document useful (0 votes)
52 views

Further Linear Algebra. Chapter VI. Inner Product Spaces

This document provides definitions and theorems regarding inner product spaces. It begins by defining positive definite symmetric bilinear forms on real vector spaces and Hermitian forms on complex vector spaces. It then defines inner product spaces and provides examples. The document proves the Cauchy-Schwarz inequality, triangle inequality, Pythagorean theorem, and describes the Gram-Schmidt process for orthogonalizing bases. It concludes by stating any vector in an inner product space with an orthonormal basis can be written as a linear combination of the basis vectors.

Uploaded by

cons the
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
52 views

Further Linear Algebra. Chapter VI. Inner Product Spaces

This document provides definitions and theorems regarding inner product spaces. It begins by defining positive definite symmetric bilinear forms on real vector spaces and Hermitian forms on complex vector spaces. It then defines inner product spaces and provides examples. The document proves the Cauchy-Schwarz inequality, triangle inequality, Pythagorean theorem, and describes the Gram-Schmidt process for orthogonalizing bases. It concludes by stating any vector in an inner product space with an orthonormal basis can be written as a linear combination of the basis vectors.

Uploaded by

cons the
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 15

Further linear algebra. Chapter VI.

Inner
product spaces.
Andrei Yafaev

1 Geometry of Inner Product Spaces


Definition 1.1 Let V be a vector space over R and let h, i be a symmetric
bilinear form on V . We shall call the form positive definite if for all non-zero
vectors v V we have
hv, vi > 0.

Notice that a symmetric bilinear form is positive definite if and only if its
canonical form (over R) is In .
Clearly x21 + . . . + x2n is positive definite on Rn . Conversely, suppose B is
a basis such that the matrix with respect to B is the canonical form. For any
basis vector bi , the diagonal entry satisfies hbi , bi i > 0 and hence hbi , bi i = 1.

Definition 1.2 Let V be a vector space over C. A Hermitian form on V is


a function h, i : V V C such that:

For all u, v, w V and all C,

hu + v, wi = hu, wi + hv, wi;

For all u, v V ,
hu, vi = hv, ui.

Example 1.1 The simplest example is the following : take V = C, then


< z, w >= z w is a hermitian form on C.
Here A is
A matrix A Mn (C) is called a Hermitian matrix if At = A.
the matrix obtained from A by applying complex conjugation to the entries.

1
If A is a Hermitian matrix then the following is a Hermitian form on Cn :

hv, wi = v t Aw.

In fact every Hermitian form on Cn is one of these.


To see why, suppose wePare given a Hermitian
P form <, >. Choose a basis
B = (b1 , . . . , bn ). Let v = i i bi and w = j j bj . We calculate
X X X
< v, w >=< i bi , j bj >= i j < bi , bj >= v t Aw
i j i,j

where A = (< bi , bj >). Of course At = A because < bi , bj >= < bj , bi >.


A matrix A satisfying At = A is called hermitian.

Example 1.2 If V = Rn , then <, > defined by < x1 , . . . , xn , y1 , . . . , yn >=


P
i,j xi yj is called the standard inner product.
If V = Cn , then <, > defined by < z1 , . . . , zn , w1 , . . . , wn >= i,j zi wj is
P
called the standard (hermitian) inner product.

Note that a Hermitian form is conjugate-linear in the second variable, i.e.


wi.
hu, v + wi = hu, vi + hu,

Note also that by the second axiom

hu, ui R.

Definition 1.3 A Hermitian form is positive definite if for all non-zero


vectors v we have
hv, vi > 0.
In other words, < v, v > 0 for all v and < v, v >= 0 if and only if
v = 0.

Clearly, the fom z w is positive definite.

Definition 1.4 By an inner product space we shall mean one of the follow-
ing:
either A finite dimensional vector space V over R with a positive definite
symmetric bilinear form;

2
or A finite dimensional vector space V over C with a positive definite
Hermitian form.

We shall often write K to mean the field R or C, depending on which is


relevant.

Example 1.3 Consider the vector space V of all continuous functions [0, 1]
C.
Then we can define
Z 1
hf, gi = f (x)g(x)dx.
0

This defines an inner product on V (easy exercise).


Another example. Let V = Mn (R) the vector space of n n-matrices with
real entries. Then
< A, B >= tr(AB t )
is an inner product on V .
t
Similarly, if V = Mn (C) and < A, B >= tr(AB ) is an inner product.

Definition 1.5 Let V be an inner product space. We define the norm of a


vector v V by p
||v|| = hv, vi.
= ||2 for for v V we have ||v|| =
Lemma 1.4 For K we have
|| ||v||.

The proof is obvious.

Theorem 1.5 (Cauchy-Schwarz inequality) If V is an inner product space


then
u, v V |hu, vi| ||u|| ||v||.

Proof. If v = 0 then the result holds so suppose v 6= 0. We have for all


K,
hu v, u vi 0.
Expanding this out we have:
vi + ||2 ||v||2 0.
||u||2 hv, ui hu,

3
hu,vi
Setting = ||v||2
we have:

hu, vi 2

2hu, vi hv, ui ||v||2 0.
||u|| hv, ui hu, vi +
||v||2 ||v||2 ||v||
2

Multiplying by ||v||2 we get

||u||2 ||v||2 2|hu, vi|2 + |hu, vi|2 0.

Hence
||u||2 ||v||2 |hu, vi|2 .
Taking the square root of both sides we get the result. 

Theorem 1.6 (Triangle inequality) If V is an inner product space with


norm || || then

u, v V ||u + v|| ||u|| + ||v||.

Proof. We have

||u + v||2 = hu + v, u + vi
= ||u||2 + 2 hu, vi + ||v||2 .

Notice that |(< u, v >)| | < u, v > | hence

||u + v||2 ||u||2 + 2| < u, v > | + ||v||2

So the CauchySchwarz inequality implies that

||u + v||2 ||u||2 + 2||u|| ||v|| + ||v||2 = (||u|| + ||v||)5 .

Hence
||u + v|| ||u|| + ||v||.


Definition 1.6 Two vectors v, w in an inner product space are called or-
thogonal if hv, wi = 0.

4
Theorem 1.7 (Pythagoras Theorem) Let (V, <, >) be an inner product
space. If v, w V are orthogonal, then

||v||2 + ||w||2 = ||v + w||2

Proof. Since

||v + w||2 = hv + w, v + wi = ||v||2 + 2hv, wi + ||w||2 ,

so we have
||v||2 + ||w||2 = ||v + w||2
if hv, wi = 0. 

2 GramSchmidt Orthogonalisation
Definition 2.1 Let V be an inner product space. We shall call a basis B of
V an orthonormal basis if hbi , bj i = i,j .

Proposition 2.1 If B is an orthonormal basis then for v, w V we have:

hv, wi = [v]tB [w]B .

Proof. If the basis B = (b1 , . . . , bn ) is orthonormal, then the matrix of <, >
in this basis is the identity In . The proposition follows. 

Theorem 2.2 (GramSchmidt Orthogonalisation) Let B be any basis.


Then the basis C defined by

c 1 = b1
hb2 , c1 i
c 2 = b2 c1
hc1 , c1 i
hb3 , c1 i hb3 , c2 i
c3 = b3 c1 c2
hc1 , c1 i hc2 , c2 i
..
.
n1
X hbn , cr i
c n = bn cr ,
r=1
hcr , cr i

5
is orthogonal. Furthermore the basis D defined by
1
dr = cr ,
||cr ||

is orthonormal.

Proof. Clearly each bi is a linear combination of C, so C spans V . As the


cardinality of C is dim V , C is a basis. It follows also that D is a basis. Well
prove by induction that {c1 , . . . , cr } is orthogonal. Clearly any one vector is
orthogonal. Suppose {c1 , . . . , cr1 } are orthogonal. The for s < r we have
r1
X hbr , ct i
hcr , cs i = hbr , cs i hct , cs i.
t=1
hct , ct i

By the inductive hypothesis we have

hbr , cs i
hcr , cs i = hbr , cs i hcs , cs i. = hbr , cs i hbr , cs i = 0.
hcs , cs i

(notice that < ct , cs >= 0 unless t = s). This shows that {c1 , . . . , cr } are
orthogonal. Hence C is an orthogonal basis. It follows easily that D is
orthonormal. 

This theorem shows in particular that an orthonormal basis always ex-


ists. Indeed, take any basis and turn it into an orthonormal one by applying
Gram-Schmidt process to it.

Proposition 2.3 If V is an inner product space with an Porthonormal basis


n
B = {b1 , . . . , bn }, then any v V can be written as v = i=1 hv, ei iei .
Pn Pn
Proof. We have v = i=1 i ei and hv, ej i = i=1 i hei , ej i = j . 

Definition 2.2 Let S be a subspace of an inner product space V . The or-


thogonal complement of S is defined to be

S = {v V : w S hv, wi = 0}.

6
Theorem 2.4 If (V, <, >) is an inner product space and W is a subspace of
V then
V = W W ,
and hence any v V can be written as

v = w + w ,

for unique w W and w W .

Proof. We show first that V = W + W .


Let E = {e1 , . . . , en } be an orthonormal basis for V , such that {e1 , . . . , er }
is a basis for W . This can be constructed by Gram-Schmidt orthogonalisa-
tion. (choose a basis {b1 , . . . , br } for W and complete to a basis {b1 , . . . , bn }
of V .
Then apply Gram-Schmidt process. Notice that in Gram-Schmidt pro-
cess, when constructing orthonormal basis, the vectors c1 , . . . , ck lie in the
space generated by c1 , . . . , ck1 , bk . It follows that the process will give an
orthonormal basis e1 , . . . , en such that e1 , . . . , er is an orthonormal basis of
W .)
If v V then
X r Xn
v= i ei + i ei .
i=1 i=r+1

Now r
X
i ei W.
i=1

If w W then there exist i R such that


r
X
w= i ei .
i=1

So * +
n
X r
X n
X
w, i ej = i j hei , ej i = 0.
j=r+1 i=1 j=r+1

Hence n
X
i ei W .
i=r+1

7
Therefore
V = W + W .
Next suppose v W W . So hv, vi = 0 and so v = 0.
Hence V = W W and so any vector v V can be expressed uniquely
as
v = w + w ,
where w W and w W . 

3 Adjoints.
Definition 3.1 An adjoint of a linear map T : V V is a linear map T
such that hT (u), vi = hu, T (v)i for all u, v V .

Theorem 3.1 (existence and uniqueness) Every T : V V has a unique


adjoint. If T is represented by A (w.r.t. an orthonormal basis) then T is
represented by At .

Proof. (Existence) Let T be the linear map represented by At . Well prove


that it is an adjoint of A.

hT v, wi = [v]t At [w] = [v]t At [w]. = hv, T wi.

Notice that here we have used that the basis is orthonormal : we said that
the matrix of <, > was the identity. (Uniqueness) Let T , T be two adjoints.
Then we have
hu, (T T )vi = 0.
for all u, v V . In particular, let u = (T T )v, then ||(T T )v|| = 0
hence T (v) = T (v) for all v V . Therefore T = T . 

Example 3.2 Consider V = C2 with the standard orthonormal basis and let
T be represented by  
1 i
A=
i 1
Then T = T (such a linear map is called autoadjoint).
Notice that T being self-adjoint is equivalent to the matrix
representing it being hermitian

8
 
2i 1+i
A=
1 + i i
Then T = T
t
We also see that T = T (using that T is represented by A ).

4 Isometries.
Theorem 4.1 If T : V V be a linear map of a Euclidean space V then
the following are equivalent.

(i) T T = Id (i.e. T = T 1 ).

(ii) u, v V hT u, T vi = hu, vi. (i.e. T preserves the inner product.)

(iii) v V ||T v|| = ||v||. (i.e. T preserves the norm.)

Definition 4.1 If T satisfies any of the above (and so all of them) then T
is called an isometry.
t
We also see that T = T (using that T is represented by A ).

Proof. (i) = (ii)


Let u, v V then
hT u, T vi = hu, T T vi = hu, vi ,
since T = T 1 .
(ii) = (iii)
If v V then
||T v||2 = hT v, T vi
so by (ii)
||T v||2 = hv, vi = ||v||2 .
Hence ||T v|| = ||v||, so (iii) holds.
(iii) = (ii) We just show that the form can be recovered from the
norm. We have

2hu, vi = ||u + v||2 ||u||2 ||v||2 , hv, wi = hv, iwi.

9
For the second equality, notice that,

2 < v, iw >=< v, iw > +< v, iw > = i < v, w > +i< v, w > =


1
i(< v, w > < v, w >) = (< v, w > < v, w >) = 2 < v, w >
i
Now suppose ||T v|| = ||v|| for all v. Take u, v V . We have ||T (u+v)|| =
||u + v||, ||T (u)|| = ||u||, and ||T (u)|| = ||v||. It follows that

2 < T (u), T (v) >= ||T (u)+T (v)||2 ||T (u)||2 ||T (v)||2 = ||u+v||2 ||u||2 ||v||2 = 2 < u, v >

and the second inequality shows that

< T (u), T (v) >= R < T (u), iT (v) >= R(< T (u), T (iv) >) = R(< u, iv >) = < u, v >

. Hence
< T (u), T (v) >=< u, v >
.
(ii) implies (i):

hT T u, vi = hT u, T vi
= hu, vi .

Therefore < (T T I)u, v >= 0 for all v. In particular, take v = (T T I)u,


then h(T T I)u, (T T 1)ui = 0. Therefore T T = I. 

Notice that in an orthonormal basis, an isometry is represented


t
by a matrix such that A = A1 .
We let On (R) be the set of n n real matrices satisfying AAt = In (in
other words At = A1 ). f An (R) then det A = 1. If A On (R) then
At = A1 so
det A = det At = det(A1 ) = det A1 .
Therefore det(A)2 = 1 and det A = 1.

Theorem 4.2 The following are equivalent.


(i) A On (R).

(ii) The columns of A form an orthonormal basis for Rn (for the standard
inner product on Rn ).

10
(iii) The rows of A form an orthonormal basis for Rn .

Proof. We prove (i) (ii) (the proof of (i) (iii) is identical).


Consider At A. If A = [C1 , . . . , Cn ], so the jth column of A is Cj , then
the (i, j)th entry of At A is Cit Cj .
So At A = In Cit Cj = i,j hCi , Cj i = i,j {C1 , . . . , Cn }
is an orthonormal basis for Rn . 

For example take the matrix:


 
1/2 1/ 2
1/ 2 1/ 2
This matrix is in O2 (R).
In fact it is the matrix of rotation by angle /4.

Theorem 4.3 Let V be a Euclidean space with orthonormal basis E = {e1 , . . . , en }.


If F = {1 , . . . ,n } is a basis for V and P is the transition matrix from E to
F, then
P n (R) F is an orthonormal basis for V .

Proof. The jth column of P is [fj ]E so


n
X
fj = pk,j ek .
k=1

Hence
* n n
+ n
n X n
X X X X
hfi , fj i = pk,i ek , pl,j el = pk,i pl,j hek , el i = pk,i pk,j = (P t P )i,j .
k=1 l=1 k=1 l=1 k=1

So F is an orthonormal basis for Rn hfi , fj i = i,j iff P t P = In


P n (R). 

Notice that it is NOT true that matrices in On (R) are diagonal-


isable.
Indeed, take  
cos() sin()
sin() cos()

11
where is not a multiple of .
The characteristic polynomial is x2 2 cos()x+1. Then, as cos()2 1 <
0, there are no real eigenvalues and the matrix is not diagonalisable.
Notice that for a given matrix A, it is easy to check that columns are
orthogonal. If that is the case, then A is in On (R) and it is easy to calculate
inverse : A1 = At .

5 Orthogonal Diagonalisation.
Definition 5.1 Let V be an inner space. A linear map T : V V is
self-adjoint if
T = T

Notice that in an othonormal basis, T is represented by a matrix A such


t
that A = A. In particular if V is real, then A is symmetric.

Theorem 5.1 If A Mn (C) is Hermitian then all the eigenvalues of A are


real.
t
Proof. Recall that Hermitian means that A = A and that this implies that
< Au, v >=< u, Av > for all u, v. Let be an eigenvalue of A and let v 6= 0
be a corresponding eigenvector. Then

Av = v

It follows that

< Av, v >= < v, v >=< v, Av >= < v, v >

As v 6= 0, we can divide by < v, v >= ||v||2 6= 0 hence we can divide by it.


It follows that = 

In particular a real symmetric matrix always has an eigenvalue : take a


complex eigenvalue (always exists !), then by the above theorem it will be
real.

Theorem 5.2 (Spectral theorem) Let T : V V be a self-adjoint linear


map of an inner product space V . Then V has an orthonormal basis of
eigenvectors.

12
Proof. This is rather similar to Theorem 5.4.
We use induction on dim(V ) = n. True for n = 1 so suppose the result
holds n 1 and let dim(V ) = n.
Since T is self-adjoint, if E is an orthonormal basis for V and A is the
matrix representing T in E then
t
A=A.

So A is Hermitian. Hence by Theorem 6.18 A has a real eigenvalue .


So there is a vector e1 V \ {0} such that T e1 = e1 . Normalizing
(dividing by ||e1 ||) we can assume that ||e1 || = 1.
Let W = Span{e1 } then by Theorem 6.9 we have V = W W . Now

n = dim(V ) = dim(W ) + dim(W ) = 1 + dim(W ),

so dim(W ) = n 1.
We claim that T : W W , i.e. T (W ) W . Let w = e1 W ,
R and v W . Then

hw, T vi = hT w, vi = hT w, vi = hT (e1 ), vi = hT e1 , vi = he1 , vi = 0,

since e1 W . Hence T : W W .
By induction there exists an orthonormal basis of eigenvectors {e2 , . . . , en }
for W . But V = W W so E = {e1 , . . . , en } is a basis for V and he1 , ei i = 0
for 2 i n and ||e1 || = 1. Hence E is an orthonormal basis of eigenvectors
for V . 

Theorem 5.3 Let T : V V be a self-adjoint linear map of a Euclidean


space V . If , are distinct eigenvalues of T then

u V v V hu, vi = 0.

Proof. If u V and v V then

hu, vi = hu, vi = hT u, vi = hu, T vi = hu, T vi = hu, vi = hu, vi .

So ( ) hu, vi = 0, with 6= . Hence hu, vi = 0. 

13
Example 5.4 Let  
1 i
A=
i 1
This matrix is self-adjoint.
One calculates the cracteristic polynomial and finds t(t 2) (in particular
the minimal polynomial is the same, hence you know that the matrix is di-
agonalisable for other reasons than being self-adjoint). For eigenvalue zero,
one finds eigenvector  
i
1
 
i
For eigenvalue 2, one finds Then we normalise the vectors : v1 =
    1
1
i i
and v1 = 12 We let
2 1 1
 
1 i i
P =
2 1 1
and  
1 0 0
P AP =
0 2

In general the procedure for orthogonal orthonormalisationis as


follows.
Let A be an n n self-adjoint matrix.
Find eigenvalues i and eigenspaces V1 (i ). Because it is diagonalisable,
you will have:
V = V1 (1 ) Vr (r )
Choose a basis for V as union of bases of V1 (i ). Apply Gram-Schmidt to it
to get an orthonormal basis.
For example :

1 2 2
A = 2 4 4
2 4 4
This matrix is symmetric hence self-adjoint.
One calculates the characteristic polynomial and finds 2 ( 9).

14

1
For V1 (9), one finds v1 = 2 To make this orthonormal, divide by
2
the norm, i.e replace v1 by 13 v1 .
For V1 (0), one finds V1 (0) = Span(v2 , v3 ) with

2
v3 = 1
0

and
2
v4 = 0
1
By Gram-Schmidt process we replace v3 by

2
1
1
5 0

and v4 by
2
1
4
3 5 5
Let
1/3 2/5 2/3 5
P = 2/3 1/ 5 4/35
2/3 0 5/3 5
We have
9 0 0
P 1 AP = 0 0 0
0 0 0

15

You might also like