Inner Product Spaces

advertisement
Outline
Section 6: Inner Product Spaces
6.1 Inner Products
6.2 Angle and orthogonality in inner vector spaces
6.3 Orthogonal Bases; Gram-Schmidt Process
6.5 Orthogonal Matrices: Change of Basis
6.1 Inner Products
Definition
An inner product on real vector space V is a function that
associates a real number hu, vi with each pair of vectors u and
v in V such that ∀u, v, w ∈ V and ∀k ∈ R:
1. hu, vi = hv, ui
2. hu + v, wi = hu, wi + hv, wi
3. hk u, vi = k hu, vi
4. hv, vi ≥ 0 and hv, vi = 0 iff v = 0
A real space with an inner product is called a real inner
product space.
Example
For u = (u1 , . . . , un ) and v = (v1 , . . . , vn ) from R n and positive
real numbers w1 , . . . , wn the formula
hu, vi = w1 u1 v1 + · · · + wn un vn
defines the weighed Euclidean inner product.
Example
Verify that for u, v ∈ R 2 the weighted Euclidean inner product
hu, vi = 3u1 v1 + 2u2 v2
satisfies the four inner product axioms.
Definition
If V is an inner product space, then the norm of u ∈ V is
kuk = hu, ui1/2
The distance between u, v ∈ V is defined by
d(u, v) = ku − vk
Example
If V = R n and u, v ∈ V then
1/2
kuk = hu, ui
1/2
= (u · u)
q
= u12 + · · · + un2
q
d(u, v) = ku − vk = (u1 − v1 )2 + · · · + (un − vn )2
Inner products generated by matrices



Let u = 

u1
u2
..
.






 and v = 


v1
v2
..
.



 be in R n and A be an

vn
un
invertible n × n matrix. Then
hu, vi = Au · Av
is the inner product on R n generated by A.
Since u · v = vT u, one has
hu, vi = (Av)T Au
or, equivalently,
hu, vi = vT AT Au
Example
The Euclidean inner product hu, vi = 3u1 v1 + 2u2 v2 is
generated by
· √
¸
3 √0
A=
2
0
Indeed,
· √
¸· √
¸·
¸
3 √0
3 √0
u1
hu, vi = [v1 v2 ]
u2
0
2
0
2
·
¸·
¸
3 0
u1
= [v1 v2 ]
0 2
u2
= 3u1 v1 + 2u2 v2 .
Example
If
·
U=
u1 u2
u3 u4
¸
·
and V =
v1 v2
v3 v4
¸
then the following formula defines an inner product on M22 :
hU, V i = tr(U T V ) = tr(V T U) = u1 v1 + u2 v2 + u3 v3 + u4 v4
Example
If
p = a0 + a1 x + a2 x 2
and
q = b0 + b1 x + b2 x 2
then the following formula defines an inner product on P2 :
hp, qi = a0 b0 + a1 b1 + a2 b2
Theorem (Properties of Inner Products)
∀u, v ∈ V and k ∈ R:
1. h0, vi = hv, 0i = 0
2. hu, k vi = k hu, vi
3. hu, v ± wi = hu, vi ± hu, wi
4. hu ± v, wi = hu, wi ± hv, wi
Proof.
For part (3):
hu, v ± wi = hv ± w, ui
(symmetry)
= hv, ui ± hw, ui
(additivity)
= hu, vi ± hu, wi
(symmetry)
6.2 Angle and orthogonality in inner vector spaces
Recall that for the dot product one has
u · v = kukkvk cos θ,
so
cos θ =
u·v
kukkvk
cos θ =
hu, vi
kukkvk
For u, v ∈ V define
Theorem (Cauchy-Schwartz Inequality)
|hu, vi| ≤ kukkvk
∀u, v ∈ V
Proof.
Without loss of generality assume u 6= 0 and v 6= 0.
Denote a = hu, ui, b = 2hu, vi, c = hv, vi. Then ∀r ∈ R:
0 ≤ hr u + v, r u + vi = hu, ui + 2hu, vi + hv, vi
= ar 2 + br + c
Hence, b2 − 4ac ≤ 0, that is,
4hu, vi2 − 4hu, uihv, vi ≤ 0
or
|hu, vi| ≤ kukkvk
Alternative forms of Cauchy-Schwartz Inequality:
hu, vi2 ≤ hu, uihv, vi and
hu, vi2 ≤ kuk2 kvk2
Theorem (Properties of Length)
∀u, v ∈ V and ∀k ∈ R:
1. kuk ≥ 0
2. kk uk = k kuk
3. kuk = 0 if and only if u = 0
4. ku + vk ≤ kuk + kvk
Theorem (Properties of Distance)
∀u, v, w ∈ V and ∀k ∈ R:
1. d(u, v) ≥ 0
2. d(u, v) = d(v, u)
3. d(u, v) = 0 if and only if u = v
4. d(u, v) ≤ d(u, w) + d(w, v)
Proof.
For the last part one has:
ku + vk2 = hu + v, u + vi
= hu, ui + 2hu, vi + hv, vi
≤ hu, ui + 2|hu, vi| + hv, vi
≤ hu, ui + 2kukkvk + hv, vi
= kuk2 + 2kukkvk + kvk2
= (kuk + kvk)2
The second inequality is based on the Cauchy-Schwartz
Inequality.
Angle between vectors
By Cauchy-Schwartz Inequality:
·
hu, vi
kukkvk
¸2
≤1
or
−1≤
hu, vi
≤1
kukkvk
Hence, there exists a unique angle θ such that:
cos θ =
hu, vi
kukkvk
and
0 ≤ θ ≤ π.
Definition
u, v ∈ V are called orthogonal if hu, vi = 0.
Example
·
U=
1 0
1 1
¸
·
and
V =
0 2
0 0
¸
are orthogonal matrices, since
hU, V i = 1 · 0 + 0 · 2 + 1 · 0 + 1 · 0 = 0
Theorem (Generalized Theorem of Pythagoras)
For any orthogonal vectors u and v in an inner product space:
ku + vk2 = kuk2 + kvk2
Proof.
ku + vk2 = hu + v, u + vi = kuk2 + 2hu, vi + kvk2
= kuk2 + kvk2
Orthogonal complements
Definition
Let W be a subspace of an inner product space V . A vector
u ∈ V is said to be orthogonal to W if it is orthogonal to every
vector in W , and the set of vectors in V that are orthogonal to
W is called the orthogonal complement of W (notation W ⊥ ).
Theorem (Properties of orthogonal complements)
If W is a subspace of a finite-dimensional inner product space
V , then
1. W ⊥ is a subspace of V
2. W ∩ W ⊥ = {0}
¡
¢⊥
3. W ⊥ = W
Proof.
For now we only prove parts (1) and (2).
1. We show that W ⊥ is closed under addition and scalar
multiplication. Indeed, ∀u, v ∈ W ⊥ and ∀w ∈ W :
hu + v, wi = hu, wi + hv, wi = 0 + 0 = 0
hk u, wi = k hu, wi = k · 0 = 0
2. If v ∈ W ∩ W ⊥ then hv, vi = 0, so v = 0.
Theorem
If A is an n × n matrix, then:
1. The nullspace of A and the row space of A are orthogonal
complements in R n w.r.t the Euclidean inner product.
2. The nullspace of AT and the column space of A are
orthogonal complements in R n w.r.t the Euclidean inner
product.
Proof.
We only prove part (1) since the one for part (2) is similar.
Assume first that v is orthogonal to every vector in the row
space of A. In particular, v is orthogonal to the row vectors
r1 , . . . , rn of A:
r1 · v = r2 · v = · · · = rn · v = 0
Since for a solution x to the linear system Ax = 0 it holds

  
r1 · x
0
 r2 · x   0 

  
 ..  =  ..  ,
 .   . 
rn · x
0
one has Av = 0, so v is in the nullspace of A.
Conversely, if Av = 0, then, by above,
r1 · v = r2 · v = · · · = rn · v = 0
Since every vector r from the row space is a l.c. of the rows,
r = c1 r1 + c2 r2 + · · · + cn rn ,
one has
r · v = (c1 r1 + c2 r2 + · · · + cn rn ) · v
= c1 (r1 · v) + c2 (r2 · v) + · · · + cn (rn · v)
= 0 + 0 + · · · + 0 = 0,
so v is orthogonal to every vector in the row space of A.
Example
Let W be the subspace of R 5 spanned by the vectors
w1 = (2, 2, −1, 0, 1), w2 = (−1, −1, 2, −3, 1),
w3 (1, 1, −2, 0, −1), w4 = (0, 0, 1, 1, 1). Find a basis for the
orthogonal complement of W .
Solution: Compose the matrix

2
2 −1
0
1
 −1 −1
2
−3
1
A=
 1
1 −2
0 −1
0
0
1
1
1




Since the vectors v1 = (−1, 1, 0, 0, 0) and v2 = (−1, 0, −1, 0, 1)
form a basis for the nullspace of A, these vectors also form a
basis for the orthogonal complement of W .
Orthogonal Bases
Definition
A set of vectors is called orthogonal set if every two distinct
vectors are orthogonal. An orthogonal set in which each vector
has norm 1 is called orthonormal.
Example
Let u1 = (0, 1, 0), u2 = (1, 0, 1), u3 = (1, 0, −1) and assume R 3
has the Euclidean inner product. Then the set S = {u1 , u2 , u3 }
is orthogonal.
√
√
Since ku1 k = 1, ku2 k = 2, ku3 k = 2, the vectors
v1 = kuu11 k = (0, 1, 0), v2 = kuu22 k = ( √1 , 0, √1 ) and
v3 =
u3
ku3 k
2
2
= ( √1 , 0, − √1 ) build an orthonormal set.
2
2
Theorem
If S = {v1 , v2 , . . . , vn } is an orthonormal basis for an inner
product space V and u ∈ V , then
u = hu, v1 iv1 + hu, v2 iv2 + · · · + hu, vn ivn
Proof.
Since S = {v1 , v2 , . . . , vn } is a basis, u is a l.c. of the basic
vectors:
u = k1 v1 + k2 v2 + · · · + kn vn .
We show that ki = hu, vi i for i = 1, 2, . . . , n. Indeed,
hu, vi i = hk1 v1 + k2 v2 + · · · + kn vn , vi i
= k1 hv1 , vi i + k2 hv2 , vi i + · · · + kn hvn , vi i
= ki hvi , vi i = ki
Example
For an orthonormal basis v1 = (0, 1, 0), v2 = (− 54 , 0, 35 ),
v3 = ( 53 , 0, 45 ) and u = (1, 1, 1) one has
hu, v1 i = 1,
so
1
hu, v2 i = − ,
5
hu, v3 i =
7
,
5
1
7
u = v1 − v2 + v3
5
5
That is,
(1, 1, 1) = (0, 1, 0) − 15 (− 45 , 0, 53 ) + 75 ( 35 , 0, 45 )
Theorem
If S is an orthonormal basis for an n-dimensional inner product
space, and if (u)S = (u1 , . . . , un ) and (v)S = (v1 , . . . , vn ) then
q
(a) kuk = u12 + u22 + · · · + un2
p
(b) d(u, v) = (u1 − v1 )2 + (u2 − v2 )2 + · · · + (un − vn )2
(c) hu, vi = u1 v1 + u2 v2 + · · · + un vn
Proof for part (a):
kuk
2
= hu, ui =
* n
X
ui vi ,
n
X
i=1
=
n X
n
X
+
uj vj
j=1
(ui uj )hvi , vj i
i=1 j=1
=
n
X
i=1
ui2 hvi , vi i =
q
u12 + u22 + · · · + un2
Theorem
If S = {v1 , v2 , . . . , vn } is an orthogonal set of non-zero vectors
in an inner product space, then S is linearly independent.
Proof.
Assume
k1 v1 + k2 v2 + · · · + kn vn = 0
We show that ki = 0, i = 1, . . . , n. For each vi ∈ S one has
hk1 v1 + k2 v2 + · · · + kn vn , vi i = h0, vi i = 0
or, equivalently,
k1 hv1 , vi i + k2 hv2 , vi i + · · · + kn hvn , vi i = 0
which for an orthogonal set reduces to
ki hvi , vi i = 0
Since vi 6= 0, we obtain ki = 0 for i = 1, 2, . . . , n.
Orthogonal Projections
Theorem (Projection Theorem)
Let W be a finite-dimensional subspace of an inner product
space V , then every vector u ∈ V can be expressed in exactly
one way as
u = w1 + w2
where w1 ∈ W and w2 ∈ W ⊥ .
We will prove this theorem later. For now:
Definition
The vector w1 in the preceeding theorem is called the
orthogonal projection of u on W and is denoted by projW u.
The vector w2 is called the component of u orthogonal to W
and is denoted by projW ⊥ u.
Theorem (Orthonormal Basis Theorem)
Let W be a finite-dimensional subspace of an inner product
space V .
(a) If {v1 , . . . , vn } is an orthonormal basis for W and u ∈ V ,
then
projW u = hu, v1 iv1 + hu, v2 iv2 + · · · + hu, vn ivn
(b) If {v1 , . . . , vn } is an orthogonal basis for W and u ∈ V ,
then
projW u =
hu, v1 i
hu, v2 i
hu, vn i
v1 +
v2 + · · · +
vn
2
2
kv1 k
kv2 k
kvn k2
Proof for part (a):
Denote
w1 = hu, v1 iv1 + hu, v2 iv2 + · · · + hu, vn ivn
and w2 = u − w1 . Then w1 ∈ W as a l.c. of its basis vectors.
We show w2 ∈ W ⊥ , that is, w2 ⊥w for any w ∈ W . Let
w = k1 v1 + k2 v2 + · · · + kn vn .
Then hw2 , wi = hu − w1 , wi = hu, wi − hw1 , wi. One has
hu, wi = hu, k1 v1 + k2 v2 + · · · + kn vn i
= k1 hu, v1 i + k2 hu, v2 i + · · · + kn hu, vn i.
Part (c) of the Orthonormal Basis Theorem implies
hw1 , wi = hu, v1 ik1 + hu, v2 ik2 + · · · + hu, vn ikn .
Hence, hw2 , wi = 0, that is, w2 ⊥w. So, w2 ∈ W ⊥ .
Theorem (Gram-Schmidt Process)
Every nonzero finite-dimensional inner product space has an
orthonormal basis.
Proof.
Given a basis {u1 , . . . , un } of V , we construct an orthonormal
one {v1 , . . . , vn }.
Step 1: Set v1 = u1 .
Step 2: Let W1 = span{v1 }. Set
v2 = u2 − projW1 u2 = u2 −
hu2 , v1 i
v1
kv1 k2
Step 3: Let W2 = span{v1 , v2 }. Set
v3 = u3 − projW2 u3 = u3 −
... and so on, (totally n steps).
hu3 , v2 i
hu3 , v1 i
v1 −
v2
2
kv1 k
kv2 k2
Example
Let u1 = (1, 1, 1), u2 = (0, 1, 1), u3 = (0, 0, 1).
Step 1: v1 = u1 = (1, 1, 1).
Step 2:
hu2 , v1 i
v1
kv1 k2
µ
¶
2
2 1 1
= (0, 1, 1) − (1, 1, 1) = − , ,
3
3 3 3
v2 = u2 −
Step 3:
hu3 , v1 i
hu3 , v2 i
v1 −
v2
2
kv1 k
kv2 k2
µ
¶
1
1/3
2 1 1
= (0, 0, 1) − (1, 1, 1) −
− , ,
3
2/3
3 3 3
µ
¶
1 1
=
0, − ,
2 2
v3 = u3 −
Proof of the uniqueness in the Projection Theorem:
The existence of an orthonormal basis for W is provided by the
Gram-Schmidt Theorem. Let
u = w1 + w2
for some w1 ∈ W and w2 ∈ W ⊥ , and assume
u = w01 + w02
for some other w01 ∈ W and w02 ∈ W ⊥ . Then
0 = (w01 − w1 ) + (w02 − w2 ),
or
w1 − w01 = w02 − w2 .
But w01 − w1 ∈ W and w02 − w2 ∈ W ⊥ . Since W ∩ W ⊥ = {0},
we get w01 = w1 and w02 = w2 .
6.5 Orthogonal Matrices: Change of Basis
Definition
A square matrix A is called orthogonal matrix if
A−1 = AT .
For an orthogonal matrix A it holds AAT = AT A = I.
Example
The following matrix is orthogonal:
 3 2
A=

AT A = 
3
7
2
7
6
7
− 67
3
7
2
7
2
7
6
7
− 37
7
− 67
2
7


3
7
6
−7
2
7
6
7
2
7
− 37
7
3
7
6
7
2
7
3
7
6
7


6
7
2
7
− 37



1 0 0
= 0 1 0 
0 0 1
Theorem
The following statements are equivalent for an n × n matrix A:
(a) A is orthogonal.
(b) The row vectors of A form an orthonormal set in R n
with the Euclidean inner product.
(c) The column vectors of A form an orthonormal set in R n
with the Euclidean inner product.
Proof for (a) ⇐⇒ (b):



AAT = 

r1 · r1 r1 · r2 · · ·
r2 · r1 r2 · r2 · · ·
..
..
..
.
.
.
rn · r1 rn · r2 · · ·
r1 · rn
r2 · rn
..
.
rn · rn


 
 
=
 
1 0 ···
0 1 ···
.. .. . .
.
. .
0 0 ···
Hence, ri · rj = 0 for i 6= j and ri · ri = 1 for i = 1, 2, . . . , n.
0
0
..
.
1





Theorem
(a) The inverse of an orthogonal matrix is orthogonal.
(b) A product of orthogonal matrices is orthogonal.
(c) If A is orthogonal then det(A) = 1 or det(A) = −1.
Proof for part (b):
If A and B are orthogonal, then
A−1 = AT
and
B −1 = B T .
One has
(AB)−1 = B −1 A−1 = B T AT = (AB)T .
Proof for the other parts is obvious.
Theorem
If A is an n × n matrix, then the following are equivalent:
(a) A is orthogonal.
(b) ∀x ∈ R n it holds kAxk = kxk.
(c) ∀x, y ∈ R n it holds Ax · Ay = x · y.
Proof.
(a) ⇒ (b): If A is orthogonal then AT A = I, so
kAxk = (Ax · Ax)1/2 = (x · AT Ax)1/2 = (x · x)1/2 = kxk.
(b) ⇒ (c): Assume kAxk = kxk. Then
Ax · Ay =
=
=
1
1
kAx + Ayk2 − kAx − Ayk2
4
4
1
1
2
kA(x + y)k − kA(x − y)k2
4
4
1
1
kx + yk2 − kx − yk2 = x · y.
4
4
.
(c) ⇒ (a): Assume Ax · Ay = x · y. Then
x · y = x · AT Ay,
which is equivalent to
x · (AT Ay − y) = 0
or
x · (AT A − I)y = 0.
In particular, for x = (AT A − I)y one gets
(AT A − I)y · (AT A − I)y = 0,
which implies
(AT A − I)y = 0.
Since the above homogeneous system is satisfied ∀y ∈ R n , we
conclude AT A − I = 0, or AT A = I, so A is orthogonal.
Change of Orthonormal Basis
Recall, that if P is a transition matrix from a basis B 0 to a basis
B, then ∀x ∈ R n :
[v]B = P[v]B 0
and
[v]B 0 = P −1 [v]B
Theorem
If P is the transition matrix from an orthonormal basis to
another orthonormal basis, then P is orthogonal.
Proof.
By the Orthonormal Basis Theorem (part (c)), ∀u ∈ R n :
kuk = k[u]B 0 k = k[u]B k = kP[u]B 0 k.
Now, let x ∈ R n and u be such that x = [u]B 0 . Then
kuk = kxk = kPxk.
Hence, P is orthogonal.
Download