You are on page 1of 16

Section 6.

1 - Inner Products and Norms


Definition. Let V be a vector space over F {R, C}. An inner product on V is a function that assigns, to
every ordered pair of vectors x and y in V , a scalar in F , denoted hx, yi, such that for all x, y, and z in V
and all c in F , the following hold:

1. hx + z, yi = hx, yi + hz, yi

2. hcx, yi = chx, yi

3. hx, yi = hy, xi, where the bar denotes complex conjugation.

4. hx, xi > 0 if x 6= 0

Note that if z is a complex number, then the statement z 0 means that z is real and non-negative.
Notice that if F = R, (3) is just hx, yi = hy, xi.
Definition. Let A Mmn (F ). We define the conjugate transpose or adjoint of A to be the n m matrix
A such that (A )i,j = Ai,j for all i, j.

Theorem 6.1. Let V be an inner product space. Then for x, y, z V and c F ,

1. hx, y + zi = hx, yi + hx, zi

2. hx, cyi = chx, yi

3. hx, 0i = h0, xi = 0

4. hx, xi = 0 x = 0

5. If hx, yi = hx, zi, for all x V , then y = z.

Proof. (Also see handout by Dan Hadley.)

1.
hx, y + zi = hy + z, xi = hy, xi + hz, xi
= hy, xi + hz, xi = hx, yi + hx, zi

2. hx, cyi = hcy, xi = chy, xi = c hy, xi = chx, yi

3. hx, 0i = hx, 0xi = 0hx, xi = h0x, xi = h0, xi

4. If x = 0 then by (3) of this theorem, hx, xi = 0. If hx, xi = 0 then by (3) of the definition, it must be
that x = 0.

5. hx, yi hx, zi = 0, for all x V . hx, y zi = 0, for all x V . Thus, hy z, y zi = 0 and we have
y z = 0 by (4). So, y = z.

p
Definition. Let V be an inner product space. For x V , we define the norm or length of x by ||x|| = hx, xi.

Theorem 6.2. Let V be an inner product space over F . Then for all x, y V and c F , the following are
true.

1. ||cx|| = |c| ||x||.

2. ||x|| = 0 if and only if x = 0. In any case ||x|| 0.

3. (Cauchy-Schwarz Inequality) | < x, y > | ||x|| ||y||.

4. (Triangle Inequality) ||x + y|| ||x|| + ||y||.

1
Proof. (Also see notes by Chris Lynd.)
1. ||cx||2 = hcx, cxi = c c hx, xi = |c|2 ||x||2 .
p p
2. ||x|| = 0 iff (hx, xi) = 0 iff hx, xi = 0 x = 0. If x 6= 0 then (hx, xi) > 0 and so (hx, xi) > 0.
3. If y = 0 the result is true. So assume y 6= 0. Finished in class.
4. Done in class.

Definition. Let V be an inner product space, x, y V are orthogonal or ( perpendicular ) if hx, yi = 0. A


subset S V is called orthogonal if x, y S, hx, yi = 0. x V is a unit vector if ||x|| = 1 and a subset
S V is orthonormal if S is orthogonal and x S, ||x|| = 1.

Section 6.2
Definition. Let V be an inner product space. Then S V is an orthonormal basis of V if it is an ordered
basis and orthonormal.
Theorem 6.3. Let V be an inner product space and S = {v1 , v2 , . . . , vk } be an orthogonal subset of V such
that i, vi 6= 0. If y Span(S), then
Xk
hy, vi i
y= vi
i=1
||vi ||2
Proof. Let
k
X
y= ai vi
i=1
Then
Xk
hy, vj i = h ai vi , vj i
i=1
k
X
= ai hvi , vj i
i=1
= aj hvj , vj i
= aj ||vj ||2

hy,vi i
So aj = ||vi || .

Corollary 1. If S also is orthonormal then


k
X
y= hy, vi ivi
i=1

Corollary 2. If S also is orthogonal and all vectors in S are non-zero then S is linearly independent.
Pk
Proof. Suppose i=1 ai vi = 0. Then for all j, aj = h0,v ii
||vi || = 0. So S is linearly independent.

Theorem 6.4. (Gram-Schmidt) Let V be an inner product space and S = {w1 , w2 , . . . , wn } V be a


linearly independent set. Define S 0 = {v1 , v2 , . . . , vn } where v1 = w1 and for 2 k n,
k1
X hwk , vi i
vk = wk vi .
i=1
||vi ||2

Then, S is an orthogonal set of non-zero vectors and Span(S 0 ) =Span(S).


0

2
Proof. Base case, n = 1. S = {w1 }, S 0 = {v1 }.
0
Let n > 1 and Sn = {w1 , w2 , . . . , wn }. For Sn1 = {w1 , w2 , . . . , wn1 } and Sn1 = {v1 , v2 , . . . , vn1 },
we have by induction that span(Sn1 ) = span(Sn1 ) and Sn1 is orthogonal. To show Sn0 is orthogonal, we
0 0

just have to show j [n 1], < vn , vj >= 0.

n1
X < wk , vk >
vn = wn vk
||vk ||2
k=1
n1
X < wk , vk >
< vn , vj > = < wn , vj > < vk , vj >
||vk ||2
k=1
n1
X < wk , vk >
= < wn , vj > < vk , vj >
||vk ||2
k=1
< wn , vj >
= < wn , vj > < vj , vj >
||vj ||2
= =0
We now show span(Sn ) = span(Sn0 ). We know dim(span(Sn )) = n, since Sn is linearly independent.
We know dim(span(Sn0 )) = n, by Corollary 2 to Theorem 6.3.
0
We know span(Sn1 ) = span(Sn1 ). So, we just need to show vn span(()Sn ).
n1
X
vn = wn aj vj
j=1
0
for some constants a1 , a2 , . . . , an1 . For all j [n 1], vj span(Sn1 ) since span(Sn1 ) = span(Sn1 ) and
0
vj Sn1 . Therefore, vn span(Sn ).
Theorem 6.5. Let V be a non-zero finite dimensional inner product space. Then V has an orthonormal
basis . Furthermore, if = {v1 , v2 , . . . , vn } and x V , then
n
X
x= hx, vi ivi .
i=1

Proof. Start with a basis of V Apply Gram Schmidt, Theorem 6.4 to get an orthogonal set 0 . Produce
from 0 by normalizing 0 . That is, multiply each x 0 by 1/||x||. By Corollary 2 to Theorem 6.3, is
linearly independent and since it has n vectors, it must be a basis of V . By Corollary 1 to Theorem 6.3, if
x V , then
Xn
x= hx, vi ivi .
i=1

Corollary 1. Let V be a finite dimensional inner product space with an orthonormal basis = {v1 , v2 , . . . , vn }.
Let T be a linear operator on V , and A = [T ] . Then for any i, j, (A)i,j = hT (vj ), vi i.
Proof. By Theorem 6.5, x V , then
n
X
x= hx, vi ivi .
i=1
So
n
X
T (vj ) = hT (vj ), vi ivi .
i=1
Hence
[T (vj )] = (hT (vj ), v1 i, hT (vj ), v2 i, . . . , hT (vj ), vn i)t
So, (A)i,j = hT (vj ), vi i.

3
Definition. Let S be a non-empty subset of an inner product space V . We define S (read S perp) to be
the set of all vectors in V that are orthogonal to every vector in S; that is, S = {x : hx, yi = 0, y S}.
S is called the orthogonal complement of S.

Theorem 6.6. Let W be a finite dimensional subspace of an inner product space V , and let y V . Then
there exist unique vectors u W , z W such that y = u + z. Furthermore, if {v1 , v2 . . . , vk } is an
orthonormal basis for W , then
X k
u= hy, vi ivi .
i=1

Proof. Let y V . Let {v1 , v2 , . . . , vk } be an orthonormal basis for W . This exists by Theorem 6.5. Let
k
X
u= hy, vi ivi .
i=1

We will show y u W .
It suffices to show that for all i [k], hy u, vi i = 0.
k
X
hy u, vi i = hy hy, vj ivj , vi i
j=1
k
X
= hy, vi i hy, vj i hvj , vi i
j=1
= hy, vi i hy, vi i
= 0

Thus y u W . Suppose x W W . Then x W and x W . So, hx, xi = 0 and we have that


x = 0.
Suppose r W and s W such that y = r + s. Then

r+s = u+z
ru = zs

This shows that both r u and z s are in both W and W . But W W = so it must be that
r u = 0 and z s = 0 which implies that r = u and z = s and we see that the representation of y is
unique.

Corollary 1. In the notation of Theorem 6.6, the vector u is the unique vector in W that is closest to
y. That is, for any x W , ||y x|| ||y u|| and we get equality in the previous inequality if and only if
x = u.
Pk
Proof. Let y V , u = i=1 < y, vi > vi . z = y u W . Let x W . Then < y u, x >= 0, since
z W .
By Exercise 6.1, number 10, if a is orthogonal to b then ||a + b||2 = ||a||2 + ||b||2 . So we have

||y x||2 = ||(u + z) x||2


= ||(u x) + z||2
= ||u x||2 + ||z||2
||z||2
= ||y u||2

Now suppose ||yx|| = ||yu||. By the squeeze theorem, ||ux||2 +||z||2 = ||z||2 and thus ||ux||2 = 0,
||u x|| = 0, u x = 0, and so u = x.

4
Theorem 6.7. Suppose that S = {v1 , v2 , . . . , vk } is an orthonormal set in an n-dimensional inner product
space V . Then
(a) S can be extended to an orthonormal basis {v1 , . . . , vk , vk+1 , . . . , vn }
(b) If W =Span(S) then S1 = {vk+1 , vk+2 , . . . , vb } is an orthonormal basis for W .
(c) If W is any subspace of V , then dim(V ) = dim(W )+ dim(W ).
Proof. (a) Extend S to an ordered basis. S 0 = {v1 , v2 , . . . , vk , wk+1 , wk+2 , . . . , wn }. Apply Gram-Schmidt
to S 0 . First k do not change. S 0 spans V . Then normalize S 0 to obtain = {v1 , . . . , vk , vk+1 , . . . , vn }. (b)

S1 = {vk+1 , vk+2 , . . . , vn } is an orthonormal
Pn set and linearly
Pn independent Also, it is a subset of W . It

must span W since if x W , x = i=1 < x, vi > vi = i=k+1 < x, vi > vi .
(c) dim(V ) = n = k + n k = dim(W ) + dim(W ).

Section 6.3:
Theorem 6.8. Let V be a finite dimensional inner product space over F , and let g : V F be a linear
functional. Then, there exists a unique vector y V such that g(x) = hx, yi, x V .
Proof. Let = {v1 , v2 , . . . , vn } be an orthonormal basis for V and let
n
X
y= g(vi )vi .
i=1

Then for 1 j n,
n
X
hvj , yi = hvj , g(vi )vi i
i=1
n
X
= g(vi )hvj , vi i
i=1
= g(vj )hvj , vj i
= g(vj )
and we have, x V, g(x) = hx, yi.
To show y is unique, suppose g(x) = hx, y0 i, x. Then, hx, yi = hx, y0 i, x. By Theorem 6.1(e), we have
y = y0 .
Example. (2b) Let V = C2 , g(z1 , z2 ) = z1 2z2 . Then V is an inner-product space with the standard inner
product h(x1 , x2 ), (y1 , y2 )i = x1 y1 + x2 y2 and g is a linear operator on V . Find a vector y V such that
g(x) = hx, yi, x V .
Sol: We need to find (y1 , y2 ) C2 such that
g(z1 , z2 ) = h(z1 , z2 ), (y1 , y2 )i, (z1 , z2 ) C2 .
That is:
z1 2z2 = z1 y 1 + z2 y 2 . (1)
Pn
Using the standard ordered basis {(1, 0), (0, 1)} for C2 , the proof of Theorem 6.8 gives that y = i=1 g(vi )vi .
So,
(y1 , y2 ) = g(1, 0)(1, 0) + g(0, 1)(0, 1)
= z1 (1, 0) + 2z2 (0, 1)
= (1, 0) 2(0, 1)
= (1, 2)
Check (1): LHS= z1 2z2 and for y1 = 1 and y 2 = 2, we have RHS= z1 2z2 .

5
Theorem 6.9. Let V be a finite dimensional inner product space and let T be a linear operator on V . There
exists a unique operator T : V V such that hT (x), yi = hx, T (y)i, x, y V . Furthermore, T is linear.
Proof. Let y V
Define g : V F as g(x) = hT (x), yi, x V .
Claim: g is linear.
g(ax + z) = hT (ax + z), yi
= haT (x) + T (z), yi
= haT (x), yi + hT (z), yi
= ag(x) + g(z).
By Theorem 6.8, there is a unique y0 V such that g(x) = hx, y0 i.
So we have hT (x), yi = hx, y0 i, x V .
We define T : V V by T (y) = y0 . So hT (x), yi = hx, T (y)i, x V
Claim: T is linear. We have for cy + z V , T (cy + z) equals the unique y0 such that hT (x), cy + zi =
hx, T (cy + z)i. But,
hT (x), cy + zi = chT (x), yi + hT (x), zi
= chx, T (y)i + hx, T (z)i
= hx, cT (y) + T (z)i
Since y0 is unique, T (cy + z) = cT (y) + T (z).
Claim: T is unique.
Let U : V V be linear such that
hT (x), yi = hx, U (y)i, x, y V
hx, T (y)i = hx, U (y)i, x, y V
= T (y) = U (y), y V
So, T = U .
Definition. T is called the adjoint of the linear operator T and is defined to be the unique operator on V
satisfying hT (x), yi = hx, T (y)i, x, y V . For A Mn (F ), we have the definition of A , the adjoint of A
from before, the conjugate transpose.
Fact. hx, T (y)i = hT (x), yi, x, y V .
Proof.
hx, T (y)i = hT (y), xi
= hy, T (x)i
= hT (x), yi

Theorem 6.10. Let V be a finite dimensional inner product space and let be an orthonormal basis for V .
If T is a linear operator on V , then
[T ] = [T ]
Proof. Let A = [T ] and B = [T ] with = {v1 , v2 , . . . , vn }.
By the corollary to Theorem 6.5,
(B)i,j = hT (vj ), vi i
= hvi , T (vj )i
= hT (vi ), vj i
= (A)j,i
= (A )i,j

6
Corollary 2. Let A be an n n matrix, then LA = (LA ) .

Proof. Use , the standard ordered basis. By Theorem 2.16,

[LA ] = A (2)
and [LA ] = A (3)

By Theorem 6.10, [LA ] = [LA ] , which equals A by (2) and [LA ] by (3).
Therefore, [LA ] = [LA ] LA = LA .

Theorem 6.11. Let V be an inner product space, T, U linear operators on V . Then

1. (T + U ) = T + U .

2. (cT ) = cT , c F .

3. (T U ) = U T (composition)

4. (T ) = T

5. I = I.

Proof. 1.

h(T + U )(x), yi = hx, (T + U ) (y)i

and

h(T + U )(x), yi = hT (x), yi + hU (x), yi


= hx, T (y)i + hx, U (y)i
= hx, (T + U )(y)i

And (T + U ) is unique, so it must equal T + u .

2.

hcT (x), yi = hx, (cT ) (y)i

and

hcT (x), yi = chT (x), yi


= chx, T (y)i
= hx, cT (y)i

3.

hT U (x), yi = hx, (T U ) (y)i

and

hT U (x), yi = hT (U (x)), yi
= hU (x), T (y)i
= hx, U (T (y))i
= hx, (U T )(y)i

7
4.

hT (x), yi = hx, (T ) (y)i

by definition and

hT (x), yi = hx, T (y)i

by Fact .

5.
hI (x), yi = hx, I(y)i = hx, yi, x, y V

Therefore, I (x) = x, x V and we have I = I.

Corollary 1. Let A and B be n n matrices. Then

1. (A + B) = A + B .

2. (cA) = cA , c F .

3. (AB) = B A

4. (A ) = A

5. I = I.

Proof. Use Theorem 6.11 and Corollary to Theorem 6.10. Or, use below.

Example. (Exercise 5b) Let A and B be m n matrices and C an n p matrix. Then

1. (A + B) = A + B .

2. (cA) = cA , c F .

3. (AC) = C A

4. (A ) = A

5. I = I.

Proof. 1.
(A + B)i,j = (A + B)j,i = (A)j,i + (B)j,i = (A)j,i + (B)j,i

and
(A + B )i,j = (A )i,j + (B )i,j = (A)j,i + (B)j,i

2. Let c F .
(cA)i,j = (cA)j,i = c(A)j,i = c(A)j,i

and
(cA )i,j = c(A )i,j = c(A)j,i

8
3.

((AC) )i,j = (AC)j,i


n
X
= (A)j,k (C)k,i
k=1
Xn
= (A)j,k (C)k,i
k=1
Xn
= (A )k,j (C )i,k
k=1
Xn
= (C )i,k (A )k,j
k=1
= (C A )i,j

Fall 2007 - The following was not covered


For x, y F n , let hx, yin denote the standard inner product of x and y in F n . Recall that if x and y are
regarded as column vectors, then hx, yin = y x.
Lemma. Let A Mmn (F ), x F n , and y F m . Then

hAx, yim = hx, A yin .

Proof.
hAx, yim = y (Ax) = (y A)x = (A y) x = hx, A yin .

Lemma. Let A Mmn (F ), x F n . Then rank(A A) =rank(A).


Proof. A A is an n n matrix. By the Dimension Theorem, rank(A A)+nullity(A A) = n. We also have,
rank(A)+nullity(A) = n.
We will show that the nullspace of A equals the nullspace of A A. We will show A Ax = 0 if and only
if Ax = 0.

0 = A Ax
0 = hA Ax, xin
0 = hAx, A xim
0 = hAx, Axim
0 = Ax

Corollary 1. If A is an m n matrix such that rank(A) = n, then A A is invertible.


Theorem 6.12. Let A Mmn (F ) and y F m . Then there exists x0 F n such that (A A)x0 = A y and
||Ax0 y|| ||Ax y|| for all x F n . Furthermore, if rank(A) = n, then x0 = (A A)1 A y.
Proof. Define W = {Ax : x F n } = R(LA ). By the corollary to Theorem 6.6, there is a unique vector
u = Ax0 in W that is closest to y. Then, ||Ax0 y|| ||Ax y|| for all x F n . Also by Theorem 6.6,
z = y u is in W . So, z = Ax0 y is in W . So,

hAx, Ax0 yim = 0, x F n .

9
By Lemma 1,
hx, A (Ax0 y)in = 0, x F n .
So, A (Ax0 y) = 0. We see that x0 is the solution for x in

A Ax = A y.

If, in addition, we know that rank(A) = n, then by Lemma 2, we have that rank(A A) = n and is therefore
invertible. So,
x0 = (A A)1 A y.

Fall 2007 - not covered until here.


Section 6.4:
Lemma. Let T be a linear operator on a finite-dimensional inner product space V . If T has an eigenvector,
then so does T .

Proof. Let v be an eigenvector of T corresponding to the eigenvalue . For all x V , we have

0 = h0, xi = h(T I)(x), xi = hv, (T I) (x)i = hv, (T I)(x)i

So v is orthogonal to (T I)(x) for all x. Thus, v 6 R(T I) and so the nullity of (T I) is not 0.
There exists x 6= 0 such that (T I)(x) = 0. Thus x is a eigenvector corresponding to the eigenvalue
of T .

Theorem 6.14. (Schur). Let T be a linear operator on a finite-dimensional inner product space V . Suppose
that the characteristic polynomial of T splits. Then there exists an orthonormal basis for V such that the
matrix [T ] is upper triangular.

Proof. The proof is by mathematical induction on the dimension n of V . The result is immediate if n = 1.
So suppose that the result is true for linear operators on (n 1)-dimensional inner product spaces whose
characteristic polynomials split. By the lemma, we can assume that T has a unit eigenvector z. Suppose
that T (z) = z and that W = span({z}). We show that W is T -invariant. If y W and x = cz W ,
then

hT (y), xi = hT (y), czi = hy, T (cz)i = hy, cT (z)i = hy, czi


= c hy, zi = c(0) = 0.

So T (y) W .
By Theorem 5.21, the characteristic polynomial of TW divides the characteristic polynomial of T and
hence splits. By Theorem 6.7(c), dim(W ) = n 1, so we may apply the induction hypothesis to TW
and obtain an orthonormal basis of W such that [TW ] is upper triangular. Clearly, = {z} is an
orthonormal basis for V such that [T ] is upper triangular.

Definition. Let V be an inner product space, and let T be a linear operator on V . We say that T is normal
if T T = T T . An n n real or complex matrix A is normal if AA = A A.

Theorem 6.15. Let V be an inner product space, and let T be a normal operator on V . Then the following
statements are true.

(a) ||T (x)|| = ||T (x)|| for all x V .

(b) T cI is normal for every c F

(c) If x is an eigenvector of T , then x is also an eigenvector of T . In fact, if T (x) = x, then T (x) = x.

10
(d) If 1 and 2 are distinct eigenvalues of T with corresponding eigenvectors x1 and x2 , then x1 and x2
are orthogonal.
Proof. (a) For any x V , we have
2
kT (x)k = hT (x), T (x)i = hT T (x), xi = hT T (x), xi
2
= hT (x), T (x)i = kT (x)k .
(b)
(T cI)(T cI) = (T cI)(T cI)
= T (T cI) cI(T cI)
= T T cT cT ccI
= T T cT cT ccI
= T T cT cT ccI
= T (T cI) cI(T cI)
= (T cI)(T cI)
= (T cI) (T cI)
(c) Suppose that T (x) = x for some x V . Let U = T I. Then U (x) = 0, and U is normal by (b).
Thus (a) implies that

0 = kU (x)k = kU (x)k = (T I)(x) = T (x) x .
Hence T (x) = x. So x is an eigenvector of T .
(d) Let 1 and 2 be distinct eigenvalues of T with corresponding eigenvectors x1 and x2 . Then using
(c), we have
1 hx1 , x2 i = h1 x1 , x2 i = hT (x1 ), x2 i = hx1 , T (x2 )i

= x1 , 2 x2 = 2 hx1 , x2 i .
Since 1 6= 2 , we conclude that hx1 , x2 i = 0.
Definition. Let T be a linear operator on an inner product space V . We say that T is self-adjoint (
Hermitian) if T = T . An n n real or complex matrix A is self-adjoint (Hermitian) if A = A .
Theorem 6.16. Let T be a linear operator on a finite-dimensional complex inner product space V . Then T
is normal if and only if there exists an orthonormal basis for V consisting of eigenvectors of T .
Proof. The characteristic polynomial of T splits over C. By Shurs Theorem there is an orthonormal basis
= {v1 , v2 , . . . , vn } such that A = [T ] is upper triangular.
() Assume T is normal. T (v1 ) = A1,1 v1 . Therefore, v1 is an eigenvector with associated eigenvalue
A1,1 .
Assume v1 , . . . , vk1 are all eigenvectors of T .
Claim: vk is also an eigenvector.
Suppose 1 , . . . , k1 are the corresponding eigenvalues. By Theorem 6.15,
T (vj ) = j vj T (vj ) = j vj .
Also, T (vk ) = A1,k v1 + A2,k v2 + + Ak,k vk .
We also know by the Corollary to Theorem 6.5 that Ai,j =< T (vj , vi >.
So,
Aj,k = < T (vk , vj >
= < vk , T (vj ) >
= < vk , j vj >
= j < v k , v j >

0 j 6= k
=
Ak,k j = k

11
So T (vk ) = Ak,k vk and we have that vk is an eigenvector of T . By induction, is a set of eigenvectors
of T .
() If is orthonormal basis of eigenvectors then T is diagonalizable by Theorem 5.1.
D = [T ] is diagonal. Hence [T ] is diagonal and equals [T ] . Diagonal matrices commute. Hence,
[T ] [T ] = [T ] [T ] .
Let x V , x = a1 v1 + a2 v2 + + an vn .
We have that T T (x) = T T (x), and hence, T T = T T .
Lemma. Let T be a self-adjoint operator on a finite-dimensional inner product space V . Then
(a) Every eigenvalue of T is real.
(b) Suppose that V is a real inner product space. Then the characteristic polynomial of T splits.
Proof. (a) Let be an eigenvalue of T . So T (x) = x for some x 6= 0. Then

x = T (x) = T (x) = (x)

( )x = 0
But x 6= 0 so = 0, and we have that = so is real.
(b) Let dim(V ) = n and be an orthonormal basis for V . Set A = [T ] . Then

A = [T ] = [T ] = [T ] = A

So A is self-adjoint.
Define TA : Cn Cn by TA (x) = Ax. Notice that [TA ] = A where is the standard ordered basis,
which is orthonormal.
So TA is self-adjoint and by (a) its eigenvalues are real.
WE know that over C the characteristic polynomial of TA factors into linear factors, t , and since
each is R, we know it also factors over R. But TA , A, and T all have the same characteristic
polynomial.

Theorem 6.17. Let T be a linear operator on a finite-dimensional real inner product space V . Then T is
self-adjoint if and only if there exists an orthonormal basis for V consisting of eigenvectors of T .
Proof. ( ) Assume T is self-adjoint. By the lemma, the characteristic polynomial of T splits. Now by
Shurs Theorem, there exists an orthonormal basis for V such that A = [T ] is upper triangular. But

A = [T ] = [T ] = [T ] = A

So A and A are both upper triangular, thus diagonal and we see that is a set of eigenvectors of T .
( ) Assume there is an orthonormal basis of V of eigenvectors of T . We know D = [T ] is diagonal with
eigenvalues on the diagonal. D is diagonal and is equal to D since it is real. But then [T ] = D = D =
[T ] = [T ] . So T = T .

Fall 2007 - The following was covered but will not be included on the final.
Section 6.5:
Definition. Let T be a linear operator on a finite-dimensional inner product space V (over F ). If ||T (x)|| =
||x|| for all x V , we call T a unitary operator if F = C and an orthogonal operator if F = R.
Theorem 6.18. Let T be a linear operator on a finite-dimensional inner product space V . Then the following
statements are equivalent.
1. T T = T T = I.

12
2. hT (x), T (y)i = hx, yi, x, y V .

3. If is an orthonormal basis for V , then T () is an orthonormal basis for V .

4. There exists an orthonormal basis for V such that T () is an orthonormal basis for V .

5. ||T (x)|| = ||x||, x V .

Proof. (1) (2)

hT (x), T (y)i = hx, T T (y)i


= hx, yi

(2) (3) Let vi , vj . Then 0 = hvi , vj i = hT (vi ), T (vj )i, so T () is orthogonal.


By Corollary 2 to Theorem 6.3, any orthogonal subset is linearly independent and since T () has n
vectors, it must be a basis of V .
Also,
1 = ||vi ||2 = hvi , vi i = hT (vi ), T (vi )i.
So, T () is an orthonormal basis of V .
(3) (4) By Graham-Schmit, there is an orthonormal basis for V . By (3), T () is orthonormal.
(4) (5) Let = {v1 , v2 , . . . , vn } be an orthonormal basis for V . Let

x = a1 v1 + a2 v2 + an vn .

Then
||x||2 = a21 hv1 , v1 i + a22 hv2 , v2 i + + a2n hvn , vn i = a21 + a22 + + a2n
And

||T (x)||2 = a21 hT (v1 ), T (v1 )i + a22 hT (v2 ), T (v2 )i + + a2n hT (vn ), T (vn )i = a21 + a22 + + a2n

Therefore, ||T (x)|| = ||x||.


(5) (1) We are given ||T (x)|| = ||x||, for all x.
We know hx, xi = 0 if and only if x = 0 and hT (x), T (x)i = 0 if and only if T (x) = 0.
Therefore, T (x) = 0 if and only if x = 0.
So N (T ) = {0} and therefore, T is invertible.
We have hx, xi = hT (x), T (x)i = hx, T T (x)i. Therefore, T T (x) = x for all x, which implies that
T T = I. But since T is invertible, it must be that T = T 1 and we have that T T = T T = I.

Lemma. Let U be a self-adjoint operator on a finite-dimensional inner product space V . If hx, U (x)i = 0
for all x V , then U = T0 . (Where T0 (x) = 0, x.)

Proof. u = U U is normal.
If F = C, by Theorem 6.16, there is an orthonormal basis of eigenvectors.
If F = R, by Theorem 6.17, there is an orthonormal basis of eigenvectors.
Let x . Then U (x) = x for some . x,

0 =< x, U (x) >=< x, x >= < x, x >

Therefore, = 0, so = 0 and U (x) = 0, x.

Corollary 1. Let T be a linear operator on a finite-dimensional real inner product space V . Then V has
an orthonormal basis of eigenvectors of T with corresponding eigenvalues of absolute value 1 if and only if
T is both self-adjoint and orthogonal.

13
Proof. () Suppose V has an orthonormal basis {v1 , v2 , . . . , vn } such that i, T (vi ) = i vi and |i | = 1.
Then by Theorem 6.17, T is self-adjoint. Well show T T = T T = I. Then by Theorem 6.18, ||T (x)|| =
||x||, x V T is orthogonal.

T T (vi ) = T (T (vi )) = T (i vi ) = i T (vi ) = i i vi = i2 vi = vi


So T T = I. Similarly, T T = I.
() Assume T is self-adjoint. Then by Theorem 6.17, V has an orthonormal basis {v1 , v2 , . . . , vn } such
that T (vi ) = i vi , i.
If T is also orthogonal, we have

|i | ||vi || = ||i vi || = ||T (vi )|| = ||vi || |i | = 1

Corollary 2. Let T be a linear operator on a finite-dimensional complex inner product space V . Then V
has an orthonormal basis of eigenvectors of T with corresponding eigenvalues of absolute value 1 if and only
if T is unitary.

Definition. A square matrix A is called an orthogonal matrix if At A = AAt = I and unitary if A A =


AA = I. We say B is unitarily equivalent to D if there exists a unitary matrix Q such that D = Q BQ.

Theorem 6.19. Let A be a complex n n matrix. Then A is normal if and only if A is unitarily equivalent
to a diagonal matrix.

Proof. () Assume A is normal. There is an orthonormal basis for F n consisting of eigenvectors of A, by


Theorem 6.16. Let = {v1 , v2 , . . . , vn }. So A is similar to a diagonal matrix D by Theorem 5.1, where the
matrix S with column i equal to vi is the invertible matrix of similarity. S 1 AS = D. S is unitary.
(Lef tarrow) Suppose A = P DP where P is unitary and D is diagonal.

AA = (P DP )(P DP ) = P DP P D P = P DD P

Also A A = P D DP , but D D = DD .

Theorem 6.20. Let A be a real nn matrix. Then A is symmetric if and only if A is orthogonally equivalent
to a real diagonal matrix.

Theorem 6.21. Let A Mn (F ) be a matrix whose characteristic polynomial splits over F .

1. If F = C, then A is unitarily equivalent to a complex upper triangular matrix.

2. If F = R, then A is orthogonally equivalent to a real upper triangular matrix.

Proof. (1) By Shurs Theorem there is an orthonormal basis = {v1 , v2 , . . . , vn } such that [LA ] = N
where N is a complex upper triangular matrix.
0
Let 0 be the standard ordered basis. Then [LA ] 0 = A. Let Q = [I] . Then N = Q1 AQ. We know
that Q is unitary since its columns are an orthonormal set of vectors and so Q Q = I.

Section 6.6
Definition. If V = W1 W2 , then a linear operator T on V is the projection on W1 along W2 if whenever
x = x1 + x2 , with x1 W1 and x2 W2 , we have T (x) = x1 . In this case,

R(T ) = W1 = {x V : T (x) = x} and N (T ) = W2 .

We refer to T as the projection. Let V be an inner product space, and let T : V V be a projection. We
say that T is an orthogonal projection if R(T ) = N (T ) and N (T ) = R(T ).

14
Theorem 6.24. Let V be an inner product space, and let T be a linear operator on V . Then T is an
orthogonal projection if and only if T has an adjoint T and T 2 = T = T .

Compare Theorem 6.24 to Theorem 6.9 where V is finite-dimensional. This is the non-finite dimensional
version.

Theorem 6.25. (The Spectral Theorem). Suppose that T is a linear operator on a finite-dimensional inner
product space V over F with the distinct eigenvalues 1 , 2 , . . . , k . Assume that T is normal if F = C and
that T is self-adjoint if F = R. For each i(1 i k), let Wi be the eigenspace of T corresponding to the
eigenvalue i , and let Ti be the orthogonal projection of V on Wi . Then the following statements are true.

(a) V = W1 W2 Wk .

(b) If Wi0 denotes the direct sum of the subspaces Wj for j 6= i, then Wi = Wi0 .

(c) Ti Tj = i,j Ti for 1 i, j k.

(d) I = T1 + T2 + + Tk .

(e) T = 1 T1 + 2 T2 + + k Tk .

Proof. Assume F = C.
(a) T is normal. By Theorem 6.16 there exists an orthonormal basis of eigenvectors of T . By Theorem 5.10,
V = W1 W2 Wk .
(b) Let x Wi and y Wj , i 6= j.
Then hx, yi = 0 and so Wi0 = Wi .
But from (1), X
dim(Wi0 ) = dim(Wj ) = dim(V ) dim(Wi ).
j6=i

By Theorem 6.7(c), we know also that

dim(Wi ) = dim(V ) dim(Wi ).

Hence Wi0 = Wi .
(c) Ti is the orthogonal projection of T on Wi . For i 6= j, x V , x = w1 + w2 + + wk , wi Wi .

Ti (x) = wi
Tj Ti (x) = Tj (wi ) = 0
Ti (wi ) = wi
Ti (Ti (x)) = Ti (x)

(d)

(T1 + T2 + + Tk )(x) = T1 (x) + T2 (x) + + Tk (x)


= w1 + w2 + + wk = x

(e) Let x = T1 (x) + T2 (x) + + Tk (x).


So, T (x) = T (T1 (x)) + T (T2 (x)) + + T (Tk (x)).
For all i, Ti (x) Wi . So

T (Ti (x)) = i Ti (x)


= 1 T1 (x) + 2 T2 (x) + + k Tk (x)
= (1 T1 + 2 T2 + + k Tk )(x).

15
Definition. The set {1 , 2 , . . . , k } of eigenvalues of T is called the spectrum of T , the sum I = T1 + T2 +
+ Tk is called the resolution of the identity operator induced by T , and the sum T = 1 T1 + 2 T2 +
+ k Tk is called the spectral decomposision of T .

Corollary 1. If F = C, then T is normal if and only if T = g(T ) for some polynomial g.


Corollary 2. If F = C, then T is unitary if and only if T is normal and || = 1 for every eigenvalue of
T.
Corollary 3. If F = C and T is normal, then T is self-adjoint if and only if every eigenvalue of T is real.
Corollary 4. Let T be as in the spectral theorem with spectral decomposition T = 1 T1 + 2 T2 + + k Tk .
Then each Tj is a polynomial in T .

16

You might also like