Anda di halaman 1dari 169

Functional Analysis (Math 920)

Lecture Notes for Spring 08


Je Schenker
Michigan State University
E-mail address: jeffrey@math.msu.edu
Contents
Comments and course information
Part 1. Hahn-Banach Theorem and Applications
Lecture 1. Linear spaces and the Hahn Banach Theorem
Lecture 2. Geometric Hahn-Banach Theorems
Lecture 3. Applications of Hahn-Banach
Part 2. Banach Spaces
Lecture 4. Normed and Banach Spaces
Lecture 5. Noncompactness of the Ball and Uniform Convexity
Lecture 6. Linear Functionals on a Banach Space
Lecture 7. Isometries of a Banach Space
Homework I
Part 3. Hilbert Spaces and Applications
Lecture 8. Scalar Products and Hilbert Spaces
Lecture 9. Riesz-Frechet and Lax-Milgram Theorems
Lecture 10. Geometry of a Hilbert space and Gram-Schmidt process
Part 4. Locally Convex Spaces
Lecture 11. Locally Convex Spaces and Spaces of Test Functions
Lecture 12. Generation of a LCS by seminorms and Frechet Spaces
Lecture 13. The dual of an LCS
Lecture 14. Spaces of distributions
Lecture 15. Applications: solving some PDEs
Lecture 16. The Dirichlet problem
Part 5. Weak Convergence and Weak Topology
iii
iv CONTENTS
Lecture 17. Dual of a Banach space
Lecture 18. Riesz-Kakutani theorem
Lecture 19. Weak convergence
Lecture 20. Weak sequential compactness, weak

convergence and the weak

topology
Lecture 21. An application: positive harmonic functions
Presentation topics
Homework II
Part 6. Convexity
Lecture 22. Convex sets in a Banach space
Lecture 23. Convex sets in a Banach space (II)
Lecture 24. Krein-Milman and Stone-Weierstrass
Lecture 25. Choquet type theorems
Part 7. Bounded Linear Maps
Lecture 26. Bounded Linear Maps
Lecture 27. Principle of Uniform Boundedness and Open Mapping Theorem
Lecture 28. The Spectrum of a Linear Map
Lecture 29. Some examples
Part 8. Compact Linear Maps
Lecture 30. Compact Maps
Lecture 31. Fredholm alternative
Lecture 32. Spectral Theory of Compact Maps
Homework III
Part 9. Compact Linear Maps in Hilbert Space
Lecture 33. Compact Symmetric Operators
Lecture 34. Min-Max
Lecture 35. Functional calculus and polar decomposition
Comments and course information
These are lecture notes for Functional Analysis (Math 920), Spring 2008. The text for
this course is Functional Analysis by Peter D. Lax, John Wiley & Sons (2002), referred to
as Lax below. In some places I follow the book closely in others additional material and
alternative proofs are given.
Other excellent texts include
M. Reed and B. Simon, Methods of Modern Mathematical Physics Vol. I: Functional
Analysis, Academic Press (1980).
W. Rudin, Functional Analysis, McGraw-Hill, 2nd ed. (1991).
(As needed, these will be referred to below as Reed and Simon and Rudin respectively.)
v
Part 1
Hahn-Banach Theorem and Applications
LECTURE 1
Linear spaces and the Hahn Banach Theorem
Reading: Chapter 1 and 3.1 of Lax
Many objects in mathematics particularly in analysis are, or may be described in
terms of, linear spaces (also called vector spaces). For example:
(1) C(`) = space of continuous functions (R or C valued) on a manifold `.
(2) (l) = space of analytic functions in a domain l C.
(3) 1
p
(j) = j integrable functions on a measure space /. j.
The key features here are the axioms of linear algebra,
Definition 1.1. A linear space A over a eld 1 (in this course 1 = R or C) is a set on
which we have dened
(1) addition: r. A r + A
and
(2) scalar multiplication: / 1. r A /r A
with the following properties
(1) (A. +) is an abelian group (+ is commutative and associative and identity and
inverses.)
identity is called 0 (zero)
inverse of r is denoted r
(2) scalar multiplication is
associative: c(/r) = (c/)r,
distributive: c(r +) = cr +/ and (c +/)r = cr +/r,
and satises 1r = r.
Remark. It follows from the axioms that 0r = 0 and r = (1)r.
Recall from linear algebra that a set of vectors o A is linearly independent if
n

j=1
c
j
r
j
= 0 with r
1
. . . . . r
n
o = c
1
= = c
n
= 0
and that the dimension of A is the cardinality of a maximal linearly independent set in A.
The dimension is also the cardinality of a minimal spanning set, where the span of a set o
is the set
span o =
_
n

j=1
c
j
r
j
: c
1
. . . . . c
n
R and r
1
. . . . . r
n
o
_
.
and o is spanning, or spans A, if span o = A.
More or less, functional analysis is linear algebra done on spaces with innite dimension.
Stated this way it may seem odd that functional analysis is part of analysis. For nite
dimensional spaces the axioms of linear algebra are very rigid: there is essentially only
1-1
1-2 1. LINEAR SPACES AND THE HAHN BANACH THEOREM
one interesting topology on a nite dimensional space and up to isomorphism there is only
one linear space of each nite dimension. In innite dimensions we shall see that topology
matters a great deal, and the topologies of interest are related to the sort of analysis that
one is trying to do.
That explains the second word in the name functional analysis. Regarding functional,
this is an archaic term for a function dened on a domain of functions. Since most of
the spaces we study are function spaces, like C(`), the functions dened on them are
functionals. Thus functional analysis. In particular, we dene a linear functional to be
a linear map / : A 1, which means
/(r +) = /(r) +/() and /(cr) = c/(r) for all r. A and c 1.
Often, one is able to dene a linear functional at rst only for a limited set of vectors
1 A. For example, one may dene the Riemann integral on 1 = C[0. 1], say, which
is a subset of the space 1[0. 1] of all bounded functions on [0. 1]. In most cases, as in the
example, the set 1 is a subspace:
Definition 1.2. A subset 1 A of a linear space is a linear subspace if it is closed
under addition and scalar multiplication:
1
.
2
1 and c 1 =
1
+c
2
1 .
For functionals dened, at rst, on a subspace of a linear space of R we have
Theorem 1.1 (Hahn (1927), Banach (1929)). Let A be a linear space over R and j a
real valued function on A with the properties
(1) j(cr) = cj(r) for all r A and c 0 (Positive homogeneity)
(2) j(r +) j(r) +j() for all r. A (subadditivity).
If / is a linear functional dened on a linear subspace of 1 and dominated by j, that is
/() j() for all 1 , then / can be extended to all of A as a linear functional dominated
by j, so /(r) j(r) for all r A.
Example. Let A = 1[0. 1] and 1 = C[0. 1]. On 1 , let /()) =
_
1
0
)(t)dt (Riemann
integral). Let j : 1 R be j()) = sup[)(r)[ : r [0. 1]. Then j satises (1) and (2)
and /()) j()). Thus we can extend / to all of 1[0. 1]. We will return to this example and
see that we can extend / so that /()) 0 whenever ) 0. This denes a nitely additive
set function on all(!) subset of [0. 1] via j(o) = /(
S
). For Borel measurable sets it turns
out the result is Lebesgue measure. That does not follow from Hahn-Banach however.
The proof of Hahn-Banach is not constructive, but relies on the following result equivalent
to the axiom of choice:
Theorem 1.2 (Zorns Lemma). Let o be a partially ordered set such that every totally
ordered subset has an upper bound. Then o has a maximal element.
To understand the statement, we need
Definition 1.3. A partially ordered set o is a set on which an order relation c / is
dened for some (but not necessarily all) pairs c. / o with the following properties
(1) transitivity: if c / and / c then c c
(2) reexivity: if c c for all c o.
(Note that (1) asserts two things: that c and c are comparable and that c c.) A subset 1
of o is totally ordered if r. 1 = r or r. An element n o is an upper bound
for 1 o if r 1 = r n. A maximal element : o satises : / = : = /.
1. LINEAR SPACES AND THE HAHN BANACH THEOREM 1-3
Proof of Hahn-Banach. To apply Zorns Lemma, we need a poset,
o = extensions of / dominated by j .
That is o consists of pairs (/
t
. 1
t
) with /
t
a linear functional dened on a subspace 1
t
1
so that
/
t
() = /(). 1 and /
t
() j(). 1
t
.
Order o as follows
(/
1
. 1
1
) (/
2
. 1
2
) 1
1
1
2
and /
2
[
Y
1
= /
1
.
If 1 is a totally ordered subset of o, let (/. 1 ) be
1 =
_
1
t
: (/
t
. 1
t
) 1
and
/() = /
t
() for 1
t
.
Since 1 is totally ordered the denition of / is unambiguous. Clearly (/. 1 ) is an upper
bound for 1. Thus by Zorns Lemma there exists a maximal element (/
+
. 1
+
).
To nish, we need to see that 1
+
= A. It suces to show that (/
t
. 1
t
) has an extension
whenever 1
t
,= A. Indeed, let r
0
A. We want /
tt
on 1
tt
= cr
0
+ : 1. c R. By
linearity we need only dene /
tt
(r
0
). The constraint is that we need
c/
tt
(r
0
) +/
t
() j(cr
0
+)
for all c. . Dividing through by [c[, since 1
t
is a subspace, we need only show
/
tt
(r
0
) j( r
0
) /
t
() (1.1)
for all 1
t
. We can nd /
tt
(r
0
) as long as
/
t
(
t
) j(
t
r
0
) j(r
0
+) /
t
() for all .
t
1
t
. (1.2)
or equivalently
/
t
(
t
+) j(r
0
+) +j(
t
r
0
) for all .
t
1
t
. (1.3)
Since
/
t
(
t
+) j(
t
+) = j(
t
r
0
+ +r
0
) j(r
0
+) +j(
t
r
0
).
(1.3), and thus (1.2), holds. So we can satisfy (1.1).
In nite dimensions, one can give a constructive proof involving only nitely many choices.
In innite dimensions the situation is a quite a bit dierent, and the approach via Zorns
lemma typically involves uncountably many choices.
LECTURE 2
Geometric Hahn-Banach Theorems
Reading: 3.2, 3.3 of Lax.
We may use Hahn-Banach to understand something of the geometry of linear spaces. We
want to understand if the following picture holds in innite dimension:
Figure 2.1. Separating a point from a convex set by a line hyperplane
Definition 2.1. A set o A is convex if for all r. o and t [0. 1] we have
tr + (1 t) o.
Definition 2.2. A point r o A is an interior point of o if for all A 0 s.t.
[t[ < = r +t o.
Remark. We can a dene a topology using this notion, letting l A be open
all r l are interior. From the standpoint of abstract linear algebra this seems to be
a natural topology on A. In practice, however, it has way too many open sets and we
work with weaker topologies that are relevant to the analysis under considerations. Much of
functional analysis centers around the interplay of dierent topologies.
We are aiming at the following
2-1
2-2 2. GEOMETRIC HAHN-BANACH THEOREMS
Theorem 2.1. Let 1 be a non-empty convex subset of A, a linear space over R, and
suppose 1 has at least one interior point. If , 1 then a linear functional / : A R s.t.
/(r) /() for all r 1. (2.1)
with strict inequality for all interior points r of 1.
This is the hyperplane separation theorem, essentially validates the picture drawn
above. A set of the form /(r) = c with / a linear functional is a hyperplane and the sets
/(r) < c are half spaces.
To accomplish the proof we will use Hahn-Banach. We need a dominating function j.
Definition 2.3. Let 1 A be convex and suppose 0 is an interior point. The gauge
of 1 (with respect to the origin) is the function j
K
: A R dened as
j
K
(r) = infc : c 0 and
r
c
1.
(Note that j
K
(r) < for all r since 0 is interior.)
Lemma 2.2. j
K
is positive homogeneous and sub-additive.
Proof. Positive homogeneity is clear (even if 1 is not convex). To prove sub-additivity
we use convexity. Consider j
K
(r +). Let c. / be such that r,c. ,/ 1. Then
t
r
c
+ (1 t)

/
1 t [0. 1].
so
r +
c +/
=
c
c +/
r
c
+
/
c +/

/
1.
Thus j
K
(r +) c +/, and optimizing over c. / we obtain the result.
Proof of hyperplane separation thm. Suces to assume 0 1 is interior and
c = 1. Let j
K
be the gauge of 1. Note that j
K
(r) 1 for r 1 and that j
K
(r) < 1 if r is
interior, as then (1 + t)r 1 for small t 0. Conversely if j
K
(r) < 1 then r is an interior
point (why?), so
j
K
(r) < 1 r 1.
Now dene /() = 1, so /(c) = c for c R. Since , 1 it is not an interior point and
so j
K
() 1. Thus j
K
(c) c for c 0 and also, trivially, for c < 0 (since j
K
0). Thus
j
K
(c) /(c)
for all c R. By Hahn-Banach, with 1 the one dimensional space c, / may be extended
to all of r so that j
K
(r) /(r) which implies (2.1).
An extension of this is the following
Theorem 2.3. Let H. ` be disjoint convex subsets of A, at least one of which has an
interior point. Then H and ` can be separate by a hyperplane /(r) = c: there is a linear
functional / and c R such that
/(n) c /()n H. `.
2. GEOMETRIC HAHN-BANACH THEOREMS 2-3
Proof. The proof rests on a trick of applying the hyperplane separation theorem with
the set
1 = H ` = n : n H and `
and the point = 0. Note that 0 , 1 since H ` = . Since 1 has an interior point
(why?), we see that there is a linear functional such that /(r) 0 for all r 1. But then
/(n) /() for all n H, `.
In many applications, one wants to consider a vector space A over C. Of course, then A
is also a vector space over R so the real Hahn-Banach theorem applies. Using this one can
show the following
Theorem 2.4 (Complex Hahn-Banach, Bohenblust and Sobczyk (1938) and Soukhom-
lino (1938)). Let A be a linear space over C and j : A [0. ) such that
(1) j(cr) = [c[j(r)c C. r A.
(2) j(r +) j(r) +j() (sub-additivity).
Let 1 be a C linear subspace of A and / : 1 C a linear functional such that
[/()[ j() (2.2)
for all 1 . Then / can be extended to all of A so that (2.2) holds for all A.
Remark. A function j that satises (1) and (2) is called a semi-norm. It is a norm if
j(r) = 0 = r = 0.
Proof. Let /
1
() = Re /(), the real part of /. Then /
1
is a real linear functional and
/
1
(i) = Re i/() = Im/(), the imaginary part of /. Thus
/() = /
1
() i/
1
(i). (2.3)
Clearly [/
1
()[ j() so by the real Hahn-Banach theorem we can extend /
1
to all of
A so that /
1
() j() for all A. Since /
1
() = /
1
() j() = j(), we have
[/
1
()[ j() for all A. Now dene the extension of / via (2.3). Given A let
= arg ln /(). Thus /() = e
i
/
1
(e
i
) (why?). So,
[/()[ = [/
1
(e
i
)[ j().

Lax gives another beautiful extension of Hahn-Banach, due to Agnew and Morse, which
involves a family of commuting linear maps. We will cover a simplied version of this next
time.
LECTURE 3
Applications of Hahn-Banach
Reading: 4.1, 4.2 of Lax.
To get an idea what one can do with the Hahn-Banach theorem lets consider a concrete
application on the linear space A = 1(o) of all real valued bounded functions on a set o.
1(o) has a natural partial order, namely r if r(:) (:) for all : o. If 0 r then r
is nonnegative. On 1(o) a positive linear functional / satises /() 0 for all 0.
Theorem 3.1. Let 1 be a linear subspace of 1(o) that contains
0
1, so
0
(:) 1
for all : o. If / is a positive linear functional on 1 then / can be extended to all of 1 as
a positive linear functional.
This theorem can be formulated in an abstract context as follows. The nonnegative
functions form a cone, where
Definition 3.1. A subset 1 A of a linear space over R is a cone if tr + : 1
whenever r. 1 and t. : 0. A linear functional on A is 1-nonnegative if /(r) 0 for
all r 1.
Theorem 3.2. Let 1 A be a cone with an interior point r
0
. If 1 is a subspace
containing r
0
on which is dened a 1 1 -positive linear functional /, then / has an extension
to A which is 1-positive.
Proof. Dene a dominating function j as follows
j(r) = inf/() : r 1. 1 .
Note that
0
tr 1 for some t 0 (since
0
is interior to 1), so
1
t

0
r 1. This
shows that j(r) is well dened. It is clear that j is positive homegeneous. To see that it is
sub-additive, let r
1
. r
2
A and let
1
.
2
be so that
j
r
j
1. Then
1
+
2
(r
1
+r
2
) 1,
so
j(r
1
+r
2
) /(
1
) +/(
2
).
Minimizing over
1,2
gives sub-additivity.
Since /(r) = /(r ) +/() /() if r 1 and r 1 we conclude that /(r) j(r)
for all r 1 . By Hahn-Banach we may extend / to all of A so that /(r) j(r) for all r.
Now let r 1. Then j(r) 0 (why?), so /(r) = /(r) 0 which shows that / is
1-positive.
The theorem on 1(o) follows from the Theorem 3.2 once we observe that the condition

0
1 implies that
0
is an interior point of the cone of positive functions. The linear
functional that one constructs in this way is monotone:
r = /(r) /().
which is in fact equivalent to positivity. This can allow us to do a little bit of analysis, even
though we havent introduced notions of topology or convergence.
3-1
3-2 3. APPLICATIONS OF HAHN-BANACH
To see how this could work, lets apply the result to the Riemann integral on C[0. 1].
We conclude the existence of a positive linear functional / : 1[0. 1] R that gives /()) =
_
1
0
)(r)dr. This gives an integral of arbitrary bounded functions, without a measurability
condition. Since the integral is a linear functional, it is nitely additive. Of course it is
not countably additive since we know from real analysis that such a thing doesnt exist.
Furthermore, the integral is not uniquely dened.
Nonetheless, the extension is also not arbitrary, and the constraint of positivity actually
pins down /()) for many functions. For example, consider
(a,b)
(the characteristic function
of an open interval). By positivity we know that
)
(a,b)
p. ). p C[0. 1] =
_
1
0
)(r)dr /(
(a,b)
)
_
1
0
p(r)dr.
Taking the sup over ) and inf over p, using properties only of the Riemann integral, we see
that /(
(a,b)
) = / c (hardly a surprising result). Likewise, we can see that /(
U
) = [l[ for
any open set, and by nite additivity /(
F
) = 1 /(1
c
) = 1 [1
c
[ = [1[ for any closed set
([ [ is Lebesgue measure). Finally if o [0. 1] and
sup /(1) : 1 closed and 1 o = inf /(l) : l open and l o
then /(
S
) must be equal to these two numbers. We see that /(
S
) = [o[ for any Lebesgue
measurable set. We have just painlessly constructed Lebesgue measure from the Riemann
integral without using any measure theory!
The rigidity of the extension is a bit surprising if we compare with what happens in
nite dimensions. For instance, consider the linear functional /(r. 0) = r dened on 1 =
(r. 0) R
2
. Let 1 be the cone (r. ) : [[ r. So / is 1 1 -positive. To extend /
to all of R
2
we need to dene /(0. 1). To keep the extension positive we need only require
/(0. 1) + 1 0
if [[ . Thus we must have [/(0. 1)[ 1,, and any choice in this interval will work.
Only when = and the cone degenerates to a half space does the condition pin /(0. 1)
down. So some interesting things happen in dimensions.
A second example application assigning a limiting value to sequences
a = (c
1
. c
2
. . . .).
Let 1 denote the space of all bounded R-valued sequences and let 1 denote the subspace of
sequences with a limit. We quickly conclude that there is an positive linear extension of the
positive linear functional lim to all of 1. We would like to conclude a little more, however.
After all, for convergent sequences,
limc
n+k
= limc
n
for any /.
To formalize this property we dene a linear map on 1 by
1(c
1
. c
2
. . . .) = (c
2
. c
3
. . . .).
(1 is the backwards shift operator or left translation.) Here,
Definition 3.2. Let A
1
. A
2
be linear spaces over a eld 1. A linear map 1 : A
1
A
2
is a function such that
1(r +c) = 1(r) +c1() r. A
1
and c 1.
3. APPLICATIONS OF HAHN-BANACH 3-3
Thus lim1a = lima for a 1. We would like the extension to have this property. Can
this be done?
The answer is given by the following a simplied version of the Theorem of Agnew and
Morse given in Chapter 3 of Lax:
Theorem 3.3. Let A be a linear space over R and let 1 : A A be a linear map. Let
j : A R be a positive homogeneous, sub-additive function invariant under 1:
j(1r) = j(r) r A.
Let / be a linear functional dened on a subspace 1 A invariant under 1,
1 1 1.
If / is invariant under 1 /(1) = /() and dominated by j then / may be extended to
all of A so that the extension is also invariant under 1 and dominated by j.
Remark. Applying this to the space 1 of bounded sequences and letting
j(a) = limsup a.
we nd that there is an extension of the linear functional lim from 1 to all of 1 such that
for all a = (c
1
. c
2
. . . .)
(1) lim
n
c
n+k
= lim
n
c
n
for all /.
(2) liminf a lima limsup a . (The rst inequality follows from the second applied
to a.)
Proof. Let us dene
(r) = inf j(r).
where the inmum is over all convex combinations of powers of 1,
=
n

j=0
c
j
1
j
(3.1)
where 1
0
= 1, the identity operator on A, 1r = r, and c
j
is any nite, non-negative sequence
with

j
c
j
= 1. Clearly (r) is positive homogeneous and (r) j(r). Also, since / and 1
are invariant under any of the form (3.1) we have
/() () 1.
To apply Hahn-Banach with as the dominating function we need to show it is sub-
additivie. To this end, let r. A and let . 1 be of the form (3.1) so that
j(r) (r) +c and j(1) () +c.
Then 1 = 1 is of the form (3.1) and
(r +) j(1(r +)) j(1r) +j(1) j(r) +j(1) (r) +() + 2c.
Thus is sub-additive.
So now we know that an extension exists to all of A with /(r) (r) for all r. However,
(r 1r) j
_
1
:
n

j=0
1
j
(1 1)r
_
=
1
:
j(r 1
n+1
r)
1
:
(j(r) +j(r)) 0.
So (r 1r) = 0 for all r and thus /(r) /(1r) = 0 for all r.
Part 2
Banach Spaces
LECTURE 4
Normed and Banach Spaces
Reading: 5.1 of Lax.
The Hahn-Banach theorem made use of a dominating function j(r). When this function
is non-negative, it can be understood roughly as a kind of distance from a point r to the
origin. For that to work, we should have j(r) 0 whenever r ,= 0. Such a function is called
a norm:
Definition 4.1. Let A be a linear space over 1 = R or C. A norm on A is a function
|| : A [0. ) such that
(1) |r| = 0 r = 0.
(2) |r +| |r| +|| (subadditivity)
(3) |cr| = [c[ |r| for all c 1 and r A (homogeneity).
A normed space is a linear space A with a norm ||.
The norm on a normed space gives a metric topology if we dene the distance between
two points via
d(r. ) = |r | .
Condition 1 guarantees that two distinct points are a nite distance apart. Sub-additivity
gives the triangle inequality. The metric is
(1) translation invariant: d(r +.. +.) = d(r. )
and
(2) homogeneous d(cr. c) = [c[d(r. ).
Thus any normed space is a metric space and we have the following notions:
(1) a sequence r
n
converges to r, r
n
r, if d(r
n
. r) = |r
n
r| 0.
(2) a set l A is open if for every r l there is a ball : | r| < c l.
(3) a set 1 A is closed if A 1 is open.
(4) a set 1 A is compact if every open cover of 1 has a nite sub-cover.
The norm denes the topology but not the other way around. Indeed two norms ||
1
and ||
2
on A are equivalent if there is c 0 such that
c |r|
1
|r|
2
c
1
|r|
2
r A.
Equivalent norms dene the same topology. (Why?)
Recall from real analysis that a metric space A is complete if every Cauchy sequence r
n
converges in A. In a normed space, a Cauchy sequence r
n
is one such that
c 0` N such that :. : ` = |r
n
r
m
| < c.
A complete normed space is called a Banach space.
4-1
4-2 4. NORMED AND BANACH SPACES
Not every normed space is complete. For example C[0. 1] with the norm
|)|
1
=
_
1
0
[)(r)[dr
fails to be complete. (It is, however, a complete space in the uniform norm, |)|
u
=
sup
x[0,1]
[)(r)[.) However, every normed space A has a completion, dened abstractly as
a set of equivalence classes of Cauchy sequences in A. This space, denoted A is a Banach
space.
Examples of Normed and Banach spaces
(1) For each j [1. ) let
/
p
= j summable sequences = (c
1
. c
2
. . . .) [

j=1
[c
j
[
p
< .
Dene a norm on /
p
via
|a|
p
=
_

j=1
[c
j
[
p
_1
p
.
Then /
p
is a Banach space.
(2) Let
/

= bounded sequences = 1(N).


with norm
|a|

= sup
j
[c
j
[. (-)
Then /

is a Banach space.
(3) Let
c
0
= sequences converging to 0 = (c
1
. c
2
. . . .) [ lim
j
c
j
= 0.
with norm (-). Then c
0
is a Banach space.
(4) Let
T = sequences with nitely many non-zero terms
= (c
1
. c
2
. . . .) [ ` `such that : ` = c
n
= 0.
Then for any j 1, T
p
= (T. ||
p
) is a normed space which is not complete. The
completion of T
p
is isomorphic to /
p
.
(5) Let 1 R
d
be a domain and let j [1. ).
(a) Let A = C
c
(1) be the space of continuous functions with compact support in
1, with the norm
|)|
p
=
__
D
[)(r)[
p
dr
_1
p
.
EXAMPLES OF NORMED AND BANACH SPACES 4-3
Then A is a normed space, which is not complete. Its completion is denoted
1
p
(1) and may be identied with the set of equivalence classes of measurable
functions ) : 1 C such that
_
D
[)(r)[
p
dr < (Lebesgue measure).
with two functions ). p called equivalent if )(r) = p(r) for almost every r.
(b) Let A denote the set of C
1
functions on 1 such that
_
D
[)(r)[
p
dr < and
_
D
[
j
)(r)[
p
dr < . , = 1. . . . . :.
Put the following norm on A,
|)|
1,p
=
_
_
D
[)(r)[
p
dr +
n

j=1
_
D
[
j
)(r)[
p
dr.
_1
p
.
Then A is a normed space which is not complete. Its completion is denoted
\
1,p
(1) and is called a Sobolev space and may be identied with the sub-
space of 1
p
(1) consisting of (equivalence classes) of functions all of whose rst
derivatives are in 1
p
(1) in the sense of distributions. (Well come back to this.)
Note that a /
1
is summable,

j
c
j
|a|
1
.
Theorem 4.1 (H olders Inqequality). Let 1 < j < and let be such that
1
j
+
1

= 1.
If a /
p
and b /
q
then
ab = (c
1
/
1
. c
2
/
2
. . . .) /
1
and

j=1
c
j
/
j

|a|
p
|b|
q
.
Proof. First note that for two non-negative numbers c. / it holds that
c/
1
j
c
p
+
1

/
q
.
For j = = 2 this is the familiar arithmetic-geometric mean inequality which follows since
(c/)
2
0. For general c. / this may be seen as follows. The function r exp(r) is convex:
exp(tr + (1 t)) t exp(r) + (1 t) exp(). (Recall from calculus that a C
2
function ) is
convex if )
tt
0.) Thus,
c/ = exp(
1
j
ln c
p
+
1

ln /
q
)
1
j
exp(ln c
p
) +
1

exp(ln /
q
) =
1
j
c
p
+
1

/
q
.
Applying this bound co-ordinate wise and summing up we nd that
|ab|
1

1
j
|a|
p
p
+
1

|b|
q
q
.
4-4 4. NORMED AND BANACH SPACES
The result follows from this bound by homogenization: we have
|a|
p
. |b|
q
= |ab| 1.
so _
_
_
_
_
a
|a|
p
b
|b|
q
_
_
_
_
_
1
1
from which the desired estimate follows by homogeneity of the norm.
Similarly, we have
Theorem 4.2 (Holders Inequality). Let 1 < j < and let be the conjugate exponent.
If ) 1
p
(1) and p 1
q
(1) then )p 1
1
(1) and
_
D
[)(r)p(r)[dr |)|
p
|p|
q
.
LECTURE 5
Noncompactness of the Ball and Uniform Convexity
Reading: 5.1 and 5.2 of Lax.
First a few more denitions:
Definition 5.1. A normed space A over 1 = R or C is called separable if it has a
countable, dense subset.
Most spaces we consider are separable, with a few notable exceptions.
(1) The space ` of all signed (or complex) measures j on [0. 1], say, with norm
|j| =
_
1
0
[j[(dr).
Here [j[ denotes the total variation of j,
[j[() = sup
Partitions A
1
, . . . , An of A
n

j=1
[j(
j
)[.
This space is a Banach space. Since the point mass

x
() =
_
1 r
0 r ,
is an element of and |
x

y
| = 2 if r ,= , we have an uncountable family of
elements of ` all at a xed distance of one another. Thus there can be no countable
dense subset. (Why?)
(2) /

is also not separable. To see this, note that to each subset of N we may
associate the sequence
A
, and
|
A

B
|

= 1
if ,= 1.
(3) /
p
is separable for 1 j < .
(4) 1
p
(1) is separable for 1 j < .
(5) 1

(1) is not separable.


Noncompactness of the Unit Ball
Theorem 5.1 (F. Riesz). Let A be a normed linear space. Then the closed unit ball
1
1
(0) = r : |A| 1 is compact if and only if A is nite dimensional.
Proof. The fact that the unit ball is compact if A is nite dimensional is the Heine-
Borel Theorem from Real Analysis.
To see the converse, we use the following
5-1
5-2 5. NONCOMPACTNESS OF THE BALL AND UNIFORM CONVEXITY
Lemma 5.2. Let 1 be a closed proper subspace of a normed space A. Then there is a
unit vector . A, |.| = 1, such that
|. |
1
2
1.
Proof of Lemma. Since 1 is proper, r A 1 . Then
inf
yY
|r | = d 0.
(This is a property of closed sets in a metric space.) We do not know the existence of a
minimizing , but we can certainly nd
0
such that
0 < |r
0
| < 2d.
Let . =
xy
0
|xy
0
|
. Then
|. | =
|r
0
|r
0
| |
|r
0
|

d
2d
=
1
2
.

Returning to the proof of the Theorem, we will use the fact that every sequence in a
compact metric space has a convergent subsequence. Thus it suces to show that if A is
innite dimensional then there is a sequence in 1
1
(0) with no convergent subsequence.
Let
1
be any unit vector and construct a sequence of unit vectors, by induction, so that
|
k
|
1
2
span
1
. . . . .
k1
.
(Note that span
1
. . . . .
k1
is nite dimensional, hence complete, and thus a closed sub-
space of A.) Since A is nite dimensional the process never stops. No subsequence of
j
can be Cauchy, much less convergent.
Uniform convexity
The following theorem may be easily shown using compactness:
Theorem 5.3. Let A be a nite dimensional linear normed space. Let 1 be a closed
convex subset of A and . any point of A. Then there is a unique point of 1 closer to . than
any other point of 1. That is there is a unique solution
0
1 to the minimization problem
|
0
.| = inf
yK
| .| . (-)
Try to prove this theorem. (The existence of a minimizer follows from compactness; the
uniqueness follows from convexity.)
The conclusion of theorem does not hold in a general innite dimensional space. Nonethe-
less there is a property which allows for the conclusion, even though compactness fails!
Definition 5.2. A normed linear space A is uniformly convex if there is a function
c : (0. ) (0. ), such that
(1) c is increasing.
(2) lim
r0
c(:) = 0.
(3)
_
_
1
2
(r +)
_
_
1 c(|r |) for all r. 1
1
(0), the unit ball of A.
UNIFORM CONVEXITY 5-3
Theorem 5.4 (Clarkson 1936). Let A be a uniformly convex Banach space, 1 a closed
convex subset of A, and . any point of A. Then the minimization problem (-) has a unique
solution
0
1.
Proof. If . 1 then
0
= . is the solution, and is clearly unique.
When . , 1, we may assume . = 0 (translating . and 1 if necessary). Let
: = inf
yK
|| .
So : 0. Now let
n
1 be a minimizing sequence, so
|
n
| :.
Now let r
n
=
n
, |
n
|, and consider
1
2
(r
n
+r
m
) =
1
2 |
n
|

n
+
1
2 |
m
|

m
=
_
1
2 |
n
|
+
1
2 |
m
|
_
(t
n
+ (1 t)
m
)
for suitable t. So t
n
+ (1 t)
m
1 and
|t
n
+ (1 t)
m
| :.
Thus
1 c(|r
n
r
m
|)
1
2
_
:
|
n
|
+
:
|
m
|
_
1.
We conclude that r
n
is a Cauchy sequence, from which it follows that
n
is Cauchy. The
limit
0
lim
n

n
exists in 1 since A is complete and 1 is closed. Clearly |
0
| = :.
Warning: Not every Banach space is uniformly convex. For example, the space C(1) of
continuous functions on a compact set 1 is not uniformly convex. It may even happen that
_
_
_
_
1
2
() +p)
_
_
_
_

= 1
for unit vectors ) and p. (They need only have disjoint support.) Lax gives an example of
a closed convex set in C[1. 1] in which the minimization problem (-) has no solution.
It can also happen that a solution exists but is not unique. For example, in C[1. 1] let
1 = functions that vanish on [1. 0]. and let ) = 1 on [1. 1]. Clearly
sup
x
[)(r) p(r)[ 1 p 1.
and the distance 1 is attained for any p 1 that satises 0 p(r) 1.
LECTURE 6
Linear Functionals on a Banach Space
Reading: 8.1 and 8.2 of Lax
Definition 6.1. A linear functional / : A 1 on a normed space A over 1 = R or C
is bounded if there is c < such that
[/(r)[ c |r| r A.
The inf over all such c is the norm |/| of /,
|/| = sup
x,=0, xX
[/(r)[
|r|
. (-)
Theorem 6.1. A linear functional / on a normed space is bounded if and only if it is
continuous.
Proof. It is useful to recall that
Theorem 6.2. Let A. 1 be metric spaces. Then ) : A 1 is continuous if and only if
)(r
n
) is a convergent sequence in 1 whenever r
n
is convergent in A.
Remark. Continuity = the sequence condition in any topological space. That the
sequence condition = continuity follows from the fact that the topology has a countable
basis at a point. (In a metric space, 1
2
n(r), say.) Specically, suppose the function is not
continuous. Then there is an open set l 1 such that )
1
(l) is not open. So there is
r )
1
(l) such that for all : 1
2
n(r) , )
1
(l). Now let r
n
be a sequence such that
(1) r
n
1
2
n(r) )
1
(l) for : odd.
(2) r
n
= r for : even.
Clearly r
n
r. However, )(r
n
) cannot converge since lim
k
)(r
2k
) = )(r) and any limit
point of )(r
2k+1
) lies in the closed set l
c
containing all the points )(r
2k+1
).
In the normed space context, note that
[/(r
n
) /(r)[ |/| |r
n
r| .
so boundedness certainly implies continuity.
Conversely, if / is unbounded then we can nd vectors r
n
so that /(r
n
) :|r
n
|. Since
this inequality is homogeneous under scaling, we may suppose that |r
n
| = 1,

:, say. Thus
r
n
0 and /(r
n
) , so / is not continuous.
The set A
t
of all bounded linear functionals on A is called the dual of A. It is a linear
space, and in fact a normed space under the norm (-). (It is straightforward to show that
(-) denes a norm.)
Theorem 6.3. The dual A
t
of a normed space A is a Banach space.
6-1
6-2 6. LINEAR FUNCTIONALS ON A BANACH SPACE
Proof. We need to show A
t
is complete. Suppose /
m
is a Cauchy Sequence. Then for
each r A we have
[/
n
(r) /
m
(r)[ |/
n
/
m
| |r| .
so /
n
(r) is a Cauchy sequence of scalars. Let
/(r) = lim
n
/
n
(r) r A.
It is easy to see that / is linear. Let us show that it is bounded. Since [ |/
n
| |/
m
| [
|/
n
/
m
| (this follows from sub-additivity), we see that the sequence |/
n
| is Cauchy, and
thus bounded. So,
[/(r)[ sup
n
|/
n
| |r|
and / is bounded. Similarly,
[/
n
(r) /(r)[ sup
mn
|/
n
/
m
| |r| .
so
|/
n
/| sup
mn
|/
n
/
m
|
and it follows that /
n
/.
Of course, all of this could be vacuous. How do we know that there are any bounded
linear functionals? Here the Hahn-Banach theorem provides the answer.
Theorem 6.4. Let
1
. . . . .
N
be ` linearly independent vectors in a normed space A
and
1
. . . . .
N
arbitrary scalars. Then there is a bounded linear functional / A
t
such that
/(
j
) =
j
. , = 1. . . . . `.
Proof. Let 1 = span
1
. . . . .
N
and dene / on 1 by
/(
N

j=1
/
j
) =
N

j=1
/
j

j
.
(We use linear independence here to guarantee that / is well dened.) Clearly, / is linear.
Furthermore, since 1 is nite dimensional / is bounded.
(Explicitly, since any two norms on a nite dimensional space are equivalent, we can nd
c such that

j
[/
j
[ |
j
| c
_
_
_
_
_

j
/
j

j
_
_
_
_
_
.
Thus, /() c sup
j
[
j
[ || for 1 .)
Thus / is a linear functional on 1 dominated by the norm ||. By the Hahn-Banach
theorem, it has an extension to A that is also dominated by ||, i.e., that is bounded.
A closed subspace 1 of a normed space A is itself a normed space. If A is a Banach
space, so is 1 . A linear functional / A
t
on A can be restricted to 1 and is still bounded.
That is there is a restriction map 1 : A
t
1
t
such that
1(/)() = /() 1.
It is clear that 1 is a linear map and that
|1(/)| |/| .
6. LINEAR FUNCTIONALS ON A BANACH SPACE 6-3
The Hahn-Banach Theorem shows that 1 is surjective. On the other hand, unless 1 = A,
the kernel of 1 is certainly non-trivial. To see this, let r be a vector in A 1 and dene /
on spanr 1 by
/(cr +) = c c 1 and 1.
Then / is bounded on the closed subspace spanr 1 and by Hahn-Banach has a closed
extension. Clearly 1(/) = 0. The kernel of 1 is denoted 1

, so
1

= / A
t
: /() = 0 1 .
and is a Banach space.
Now the quotient space A,1 is dened to be the set of cosets of Y,
1 +r : r A.
The coset 1 +r is denoted [r]. The choice of label r is, of course, not unique as r + with
1 would do just as well. It is a standard fact that
[r
1
] + [r
2
] = [r
1
+r
2
]
gives a well dened addition on 1,A so that it is a linear space.
Lemma 6.5. If 1 is a closed subspace of a normed space then
|[r]| = inf
yY
|r +|
is a norm on A,1 . If A is a Banach space, so is A,1 .
Proof. Exercise.
A bounded linear functional that vanishes on 1 , that is an element / 1

, can be
understood as a linear functional on A,1 since the denition
/([r]) = /(r)
is unambiguous. Thus we have a map J : 1

(A,1 )
t
dened by J(/)([r]) = /(r).
Conversely, there is a bounded linear map : A A,1 given by (r) = [r] and any
linear functional / (A,1 )
t
pulls back to a bounded linear functional / in 1

. Clearly
J(/ ) = /. Thus we have, loosely, that
(A,1 )
t
= 1

.
LECTURE 7
Isometries of a Banach Space
Reading: 5.3 of Lax.
Definition 7.1. Let A. 1 be normed spaces. A linear map 1 : A 1 is bounded if
there is c 0 such that
|1(r)| c |r| .
The norm of 1 is the smallest such c, that is
|1| = sup
x,=0
|1(r)|
|r|
.
Theorem 7.1. A linear map 1 : A 1 between normed spaces A and 1 is continuous
if and only if it is bounded.
Remark. The proof is a simple extension of the corresponding result for linear function-
als.
An isometry of normed spaces A and 1 is a map ` : A 1 such that
(1) ` is surjective.
(2) |`(r) `()| = |r |.
Clearly translations 1
u
: A A, 1
u
(r) = r + n are isometries of a normed linear space. A
linear map 1 : A 1 is an isometry if 1 is surjective and
|1(r)| = |r| r A.
A map ` : A 1 is ane if `(r) `(0) is linear. So, ` is ane if it is the
composition of a linear map and a translation.
Theorem 7.2 (Mazur and Ulam 1932). Let A and 1 be normed spaces over R. Any
isometry ` : A 1 is an ane map.
Remark. The theorem conclusion does not hold for normed spaces over C. In that
context any isometry is a real -ane map (`(r) `(0) is real linear), but not necessarily
a complex-ane map. For example on C([0. 1]. C) the map ) ) (complex conjugation) is
an isometry and is not complex linear.
Proof. It suces to show `(0) = 0 = ` is linear. To prove linearity it suces to
show
`
_
1
2
(r +)
_
=
1
2
(`(r) +`()) r. A.
(Why?)
Let r and be points in A and . =
1
2
(r +). Note that
|r .| = | .| =
1
2
|r | .
7-1
7-2 7. ISOMETRIES OF A BANACH SPACE
so . is half-way between r and . Let
r
t
= `(r).
t
= `(). .
t
= `(.).
We need to show 2.
t
= r
t
+
t
. Since ` is an isometry,
|r
t
.
t
| = |
t
.
t
| =
1
2
|r
t

t
| .
and all of these are equal to
1
2
|r |. So .
t
is half-way between r
t
and
t
. It may happen
that
1
2
(r
t
+
t
) is the unique point of 1 with this property (in which case we are done). this
happens, for instance, if the norm in 1 is strictly sub-additive, meaning
r
t
,=
t
= |r
t
+
t
| < |r
t
| +|
t
| .
In general, however, there may be a number of points half-way between r
t
and
t
.
So, let

1
= n A : |r n| = | n| =
1
2
|r |.
and

t
1
= n
t
1 : |r
t
n
t
| = |
t
n
t
| =
1
2
|r
t

t
|.
Since ` is an isometry, we have
t
1
= `(
1
). Let d
1
denote the diameter of
1
,
d
1
= sup
u,vA
1
|n | .
This is also the diameter of
t
1
. Now, let

2
=
_
n
1
:
1
= |n |
1
2
d
1
_
.
the set of centers of
1
. Note that .
2
since if n
1
then 2. n
1
:
|r (2. n)| = |n | = |r n| = | (2. n)| .
Similarly, let

t
2
=
_
n
t

t
1
:
t

t
1
= |n
t

t
|
1
2
d
1
_
.
Again, since ` is an isometry we have
t
2
= `(
2
). In a similar way, dene decreasing
sequences of sets,
j
and
t
j
, inductively by

j
= n
j1
:
j1
= |n |
1
2
diam(
j1
).
and

t
j
= n
t

t
j1
:
t

t
j1
= |n
t

t
|
1
2
diam(
t
j1
).
Again `(
j
) =
t
j
and .
j
since
j1
is invariant under inversion around .: n

j1
= 2. n
j1
. Since diam(
j
) 2
1j
d
1
we conclude that

j=1
= .. and

j=1

t
j
=
_
1
2
(r
t
+
t
)
_
.
Since .
t

t
j
for all ,, (-) follows.
Homework I
Homework I
Due: February 15, 2008
(1) (Ex. 1, Ch.1) Let A be a linear space. Let o

: be a collection of
subspaces. (Here is some index set, not necessarily nite or countable.) Prove
that
(a) The sum

= r
1
+ + r
n
: r
j
o

j
.
j
. , = 1. . . . . :. : N
is a subspace.
(b) The intersection

is a subspace.
(c) The union

is a subspace provided is a totally ordered set and


= o

.
(2) (Ex. 2, Ch. 4) The Cesaro means of a bounded sequence c
n
are the numbers
`
n
(c) =
1
:
n

j=1
c
j
.
A bounded sequence c is called Cesaro summable if lim
n
`
n
(c) exists.
(a) Show that any convergent sequence c is Cesaro summable and
lim
n
`
n
(c) = lim
n
c
n
.
(b) Give an explicit example of a bounded Cesaro summable sequence that is not
convergent.
(c) Show there is a Banach limit LIM such that for any bounded Cesaro summable
sequence LIM
n
c
n
= lim
n
`
N
(c). (A Banach limit is a linear functional
on the space /

of bounded sequences which agrees with the limit on convergent


sequence, is invariant under shifts, and is bounded above and below by liminf
and limsup.)
(3) (Ex. 2, Ch.5) Let A be a normed linear space and 1 a subspace of A. Prove that
the closure of 1 is a linear subspace of A.
(4) (Ex. 3, Ch.5) Show that if A is a Banach space and 1 is a closed subspace of A,
then the quotient space A,1 is complete.
(5) Let ` : A 1 be a continuous map between normed spacse A and 1 such that
`(0) = 0 and `
_
1
2
(r +)
_
=
1
2
`(r) +
1
2
`(). r. A.
Show that ` is linear.
(6) Let / c

0
be a bounded linear functional on c
0
. Prove that there is a sequence
b /
1
such that
/(a) =

j=1
/
j
c
j
.
Conclude that c

0
is isometrically isomorphic to /
1
.
(7) Let / /

1
be a bounded linear functional on /
1
. Prove that there is a sequence
b /

such that
/(a) =

j=1
/
j
c
j
.
Conclude that /

1
is isometrically isomorphic to /

.
(8) Show that /

p
is isometrically isomorphic to /
q
, 1 < j < and 1,j + 1, = 1.
(9) Let / = a /
p
: c
2n
= 0. : /
p
. Show that / is a closed subspace and that
/
p
,/ is isometrically isomorphic to /
p
.
(10) Let A
1
. A
2
be Banach spaces, with norms ||
1
. ||
2
. Dene the direct sum
A
1
A
2
= (r. ) : r A
1
and A
2
.
with coordinatewise addition and scalar multiplication. For each j [1. ] dene
the norm
|(r. )|
p
=
_
(|r|
p
1
+||
p
2
)
1/p
1 j < .
max|r|
1
. ||
2
j = .
Show that A
1
A
2
is a Banach space under ||
p
and all these norms are equivalent.
(11) Consider A = 1
1
(R) + 1
2
(R) = p + / : p 1
1
(R) and / 1
2
(R). Given
j [1. ), dene
|)|
inf
p
= inf(|p|
p
L
1
+|/|
p
L
2
)
1/p
: ) = p +/. p 1
1
and / 1
2
.
Show that |)|
inf
p
is a norm and that all of these norms are equivalent. (Challenge:
Is A a Banach space?)
(12) Consider 1 = 1
1
(R) 1
2
(R), which is a subspace of A from the previous problem.
(a) For each j [1. ), dene
|)|
+
p
=
1
2
_
|)|
p
L
1
+|)|
p
L
2
_
1/p
.
for ) 1 . Show that ||
+
p
is a norm and all of these norms are equivalent.
(b) Show that 1 is a Banach space.
(c) Show that 1 is dense in A under any of the norms ||
inf
p
.
(d) Show that |)|
inf
p
|)|
+
p
. Is there a constant c such that |)|
+
p
c |)|
inf
p
?
Part 3
Hilbert Spaces and Applications
LECTURE 8
Scalar Products and Hilbert Spaces
Reading: 6.1 and 6.2 of Lax.
Definition 8.1. A scalar product on a linear space A over R is a real valued function
. ) : A A R with the following properties
(1) Bilinearity: r r. ) and r. ) are linear functions.
(2) Symmetry: r. ) = . r).
(3) Positivity: r. r) 0 if r ,= 0. (Note that 0. 0) = 0 by bilinearity.
A (complex) scalar product on a linear space A over C is a complex valued function
. ) : A A C with the properties
(1) Sesquilinearity: r r. ) is linear and r. ) is skewlinear,
r. +c
t
) = r. ) +c

r.
t
) .
(2) Skew symmetry: r. ) = . r)

.
(3) Positivity: r. r) 0 for r ,= 0.
Given a (real or complex) scalar product, the associated norm is
|r| =
_
r. r).
Remark. A complex linear space is also a real linear space, and associated to any
complex inner product is a real inner product:
(r. ) = Re r. ) .
Note that the associated norms are the same, so the metric space structure is the same
whether or not we consider the space as real or complex. Note that
(ir. ) = (r. i). (-)
and the real and complex inner products are related by
r. ) = (r. ) i(ir. ) = (r. ) + i(r. i). (--)
Conversely, given any real inner product on a complex linear space which satises (-), (--)
gives a complex inner product.
We have not shown that the denition |r| =
_
r. r) actually gives a norm. Homogeneity
and positivity are clear. To verify subadditivity we need the following important Theorem.
Theorem 8.1 (Cauchy-Schwarz). A real or complex scalar product satises
[r. )[ |r| || .
with equality only if cr = /.
8-1
8-2 8. SCALAR PRODUCTS AND HILBERT SPACES
Remark. A corollary is that
|r| = max
|u|=1
[ r. n) [.
from which follows sub-additivity
|r +| |r| +|| .
Proof. It suces to consider the real case, since given r. we can always nd so that

e
i
r.
_
= e
i
r. ) is real. Also, we may assume ,= 0.
So let . ) be a real inner product and t R. Then
|r +t|
2
= |r|
2
+ 2t r. ) +t
2
||
2
.
Minimizing the r.h.s. over t we nd that,
t
min
=
r. )
||
2
.
and
0 |r|
2

r. )
2
||
2
.
The Cauchy-Schwarz inequality follows.
Another important, related result, is the parallelogram identity
|r +|
2
+|r |
2
= 2 |r|
2
+ 2 ||
2
.
von Neumann has shown that any norm which satises the parallelogram law comes from
an inner product.
Definition 8.2. A linear space with a scalar product that is complete in the induced
norm is a Hilbert space.
Any scalar product space can be completed in the norm. It follows from the Schwarz
inequality that the scalar product is cts. in each of its factors and extends uniquely to the
completion, which is thus a Hilbert space.
Examples:
(1) /
2
is a Hilbert space with the inner product
c. /) =

j
c
j
/

j
.
which is nite by the Holder inequality.
(2) C[0. 1] is an inner product space with respect to the inner product
r. ) =
_
1
0
r(t)(t)

dt.
It is not complete. The completion is known as 1
2
[0. 1] and can be associated with
the set of equivalence classes of Lebesgue square integrable functions.
Remark. There is no standard as to which factor of the inner product is skew-linear.
In the physics literature, it is usually the rst factor; in math it is usually the second.
8. SCALAR PRODUCTS AND HILBERT SPACES 8-3
Definition 8.3. Two vectors in an inner product space are orthogonal if
r. ) = 0.
The orthogonal complement of a set : is
o

= : . ) = 0 1 .
Lemma 8.2. Given any set o in a Hilbert space, o

is a closed subspace.
Proof. That o

is a subspace is clear. That it is closed follows, from continuity of the


inner product in each factor, since if
n
,
n
o

, then
. ) = lim
n

n
. ) = 0 for o.

Theorem 8.3. Let H be a Hilbert space, 1 a closed subspace of H and 1

the orthogonal
complement of 1 . Then
(1) Any vector r H can be written uniquely as a linear combination
r = +. 1 and 1

.
(2) (1

= 1.
To prove this theorem, we need
Lemma 8.4. Given a nonempty closed, convex subset 1 of a Hilbert space, and a point
r H, there is a unique point in 1 that is closer to r than any other point of1.
Proof. This follows if we show that H is uniformly convex, by the Theorem of Clarkson
from lecture 5. Let r. be unit vectors. It follows from the parallelogram law that
_
_
_
_
1
2
r +
_
_
_
_
2
= 1
1
4
|r |
2
.
so
_
_
_
_
1
2
r +
_
_
_
_
1
_
1
_
1
1
4
|r |
2
_
. .
|xy|
.

Proof of Theorem. According to the Lemma there is a unique point 1 closest


to a given point r H. Let = r . We claim that .
t
) = 0 for any
t
1 . Indeed, we
must have
||
2
| +t
t
|
2
= ||
2
+ 2t Re .
t
) +t
2
|
t
|
2
for any t. In other words the function
0 2t Re .
t
) +t
2
|
t
|
2
for all t.
which can occur only if Re .
t
) = 0. Since this holds for all
t
1 we get .
t
) = 0 by
complex linearity.
Thus the decomposition r = + is possible. Is it unique? Suppose r = + =
t
+
t
.
Then
t
=
t
1 1

. But . 1 1

= .. .) = 0 so . = 0.
Part (3) is left as a simple exercise.
LECTURE 9
Riesz-Frechet and Lax-Milgram Theorems
Reading: 6.3 of Lax.
We have already seen that for xed H, a Hilbert space, the map /
y
(r) = r. ) is a
bounded linear functional boundedness follows from Cauchy-Schwarz. In fancy language
/
y
embeds H into H

, the dual of H. In fact, since


|/
y
| = sup
x
[r. )[
|r|
= || .
again by Cauchy-Schwarz, this map is an isometry onto its range. In a real Hilbert space,
this is a linear map; in a complex Hilbert space, it is skew-linear:
/
y+y
= /
y
+

/
y
.
The question now comes up whether we get every linear functional in H

this way? The


answer turns out to be yes.
Theorem 9.1 (Riesz-Frechet). Let /(r) be a bounded linear functional on a Hilbert space
H. Then there is a unique H such that
/(r) = r. ) .
Before turning to the proof, let us state several basic facts, whose proof is left as an
exercise:
Lemma 9.2.
(1) Let A be a linear space and / a non-zero linear functional on A. Then the null
space of / is a linear subspace of co-dimension 1. That is, if1 = : /() = 0
then there exists r
0
, 1 and any vector r A may be written uniquely as
r = r
0
+. 1 and 1.
(2) If two linear functionals /, : share the same null space, they are constant multiples
of each other: / = c:.
(3) If A is a Banach space and / is bounded, then the null-space of / is closed.
Proof. If / = 0 then = 0 will do, and this is the unique such point .
If / ,= 0, then it has a null space 1 , which by the lemma is a closed subspace of co-
dimension 1. The orthogonal complement 1

must be one dimensional. Let be a unit


vector in 1

it is unique up to a scalar multiple. Then :(r) = r. ) is a linear functional,


with null-space 1 . Thus / = : and we may take =

.
To see that is unique, note that if r. ) = r.
t
) for all r then |
t
| = 0, so
=
t
.
In applications, one is often given not a linear functional, but a quadratic form:
9-1
9-2 9. RIESZ-FRECHET AND LAX-MILGRAM THEOREMS
Definition 9.1. Let H be a Hilbert space over R. A bi-linear form on H is a function
1 : H H R such that
r 1(r. ) and 1(r. )
are linear maps. A skew-linear form on a Hilbert space H over C is a map 1 : H H C
such that
r 1(r. ) is linear, and 1(r. ) is skew-linear.
A quadratic form refers to a bi-linear form or a skew-linear form depending on whether the
eld of scalars is R or C. A quadratic form 1 on H is bounded if there is a constant c 0
such that
[1(r. )[ c |r| || .
and is bounded from below if there is a constant / 0 such that
[1(. )[ / ||
2
.
Theorem 9.3 (Lax-Milgram). Let H be a Hilbert space, over R or C, and let 1 be a
bounded quadratic form on H that is bounded from below. Then ever bounded linear functional
/ H

may be written
/(r) = 1(r. ). for unique H.
Proof. For xed , r 1(r. ) is a bounded linear functional. By Riesz-Frechet there
exists . : H H such that
1(r. ) = r. .()) .
It is easy to see that the map .() is linear. Thus the range of .,
ran . = .() : H .
is a linear subspace of H.
Let us prove that ran . is a closed subspace. Here we need the fact that 1 is bounded
from below. Indeed,
1(. ) = . .()) .
so
/ ||
2
|| |.()| = / || |.()| .
If
n
is any sequence then
|
n

m
| /
1
|.(
n
) .(
m
)| .
Thus .(
n
) .
0
=
n
Cauchy =
n

0
, and it is easy to see we must have
.
0
= .(
0
). Thus .
0
ran ., so ran . is closed.
Now we show that ran . = H. Since ran . is closed it suces to show ran .

= 0. Let
r ran .. It follows that
1(r. ) = r. .()) = 0 H.
Thus 1(r. r) = 0 and so r = 0 since |r|
2
/
1
[1(r. r)[ .
Since ran . = H we see by Riesz-Frechet that any linear functional / may be written
/(r) = r. .()) = 1(r. ) for some . Uniqueness of follows as above, since if 1(r. ) =
1(r.
t
) for all r we conclude that |
t
| = 0 since 1 is bounded from below.
APPLICATION: RADON-NIKODYM THEOREM 9-3
Application: Radon-Nikodym Theorem
Theorem 9.4 (Radon-Nikodym). Let `, be a measurable space on which we have
dened two nite non-negative measures j and . If j() = 0 = () = 0 for all
then there exists a measurable function / : ` [0. ) such that
() =
_
A
/dj. .
Proof. Consider the measure j + and the Hilbert space 1
2
(j +),
|)|
2
=
_
M
)
2
d(j +).
Dene a linear functional
/()) =
_
M
)dj.
Since
[/())[
2
j(`)
_
M
)
2
dj j(`) |)|
2
.
it follows that / is bounded (on 1
2
(j +)). Thus p 1
2
(j +) such that
_
M
)dj =
_
M
)pd(j +).
Rewrite this as,
_
M
)(1 p)dj =
_
M
)d. (-)
Let 1 = r ` : p(r) 0 and plug ) =
F
into (-). Then
j(1)
_
F
(1 p)dj =
_
F
pd 0.
Thus j(1) = 0. Likewise let G = r ` : p(r) 1 and plug ) =
G
into (-). If
j(G) 0 then
0
_
G
(1 p)dj =
_
G
pd (G).
which is a contradiction. Thus j(G) = 0. Hence, after modifying p on the j null set 1 G
we may assume that (-) holds with
0 < p(r) 1.
But then given 1 , plugging ) = p
1

E
into (-) we get
(1) =
_
M
)pd =
_
M

E
1 p
p
dj =
_
E
1 p
p
dj.
Thus take / = (1 p),p.
LECTURE 10
Geometry of a Hilbert space and Gram-Schmidt process
Reading: 6.4 of Lax.
Recall that the linear span of a set o in a linear space A is the collection of nite linear
combinations of elements of o:
span o =
_
n

j=1

j
r
j
: r
j
o.
j
1. , = 1. . . . . :. : N
_
.
This is also the smallest subspace containing o:
span o = 1 : 1 A is a subspace and o 1 .
If A is a Banach space, it is natural to look at the smallest closed subspace containing o:
span o = 1 : 1 A is a closed subspace and o 1 .
Proposition 10.1. Let A be a Banach space. Then span o = span o.
The proof is left as an exercise.
In a Hilbert space we have a geometric characterization of span o:
Theorem 10.2. Let o H be any subset of a Hilbert space H. Then
span o =
_
o

.
That is, span o if and only if is perpendicular to everything that is perpendicular to o:
. .) = 0 for all . such that r. .) = 0 for all r o.
Proof. Recall that a closed subspace 1 satises (1

= 1 . Thus it suces to show


(span o)

= o

. Since o span o we clearly have (span o)

. On the other hand, if . o

.
Thus . is perpendicular to span o and by continuity of the scalar product . span o =
span o. Thus o

(span o)

.
Definition 10.1. A collection of vectors o in an inner product space H is called
orthonormal if
r. ) =
_
1 r = o
0 r ,= . r. o.
An orthonormal collection o is called an orthonormal basis if span o = H.
Lemma 10.3. Let o be an orthonormal set of vectors in a Hilbert space H. Then the
span o consists of all vectors of the form
r =

j=1

j
r
j
. r
j
o. , = 1. . . . . . (-)
10-1
10-2 10. GEOMETRY OF A HILBERT SPACE AND GRAM-SCHMIDT PROCESS
where the
j
are square summable:

j=1
[
j
[
2
< .
The sum converges in the Hilbert space:
_
_
_
_
_
r
n

j=1

j
r
j
_
_
_
_
_
0.
and
|r|
2
=

j=1
[
j
[
2
.
Furthermore, the sum may be written
r =

yS
r. ) .
In particular, r. ) , = 0 for only countably many elements o.
Remark. Most orthonormal sets encountered in practice are countable, so we would
tend to write o = r
1
. . . . . and
r =

j=1
r. r
j
) r
j
.
However, the lemma holds even for uncountable orthonormal sets.
Proof. It is clear that all vectors of the form (-) are in span o = span o. Furthermore
vectors of this form make up a subspace, which is easily seen to be closed. (Exercise: show
that this subspace is closed. This rests on the fact that a subset of a complete metric space
is closed i it is sequentially complete.) By denition span o is contained in this subspace.
Thus the two subspaces are equal.
The remaining formulae are easy consequences of the form (-).
Theorem 10.4. Every Hilbert space contains an orthonormal basis.
Proof. We use Zorns Lemma. Consider the collection of all orthonormal sets, with
o 1 i o 1. This collection is non-empty since any unit vector makes up a one element
orthonormal set.
A totally ordered collection has an upper bound the union of all sets in the collection.
Thus there is a maximal orthonormal set. Call it o
max
.
Suppose span o
max
A. Then, span o

max
is a non-trivial closed subspace. Let
span o

max
be a unit vector. So o
max
is an orthonormal set contradicting the fact that
o
max
is maximal.
Corollary 10.5 (Bessels inequality). Let o be any orthonormal set in a Hilbert space
H (not necessarily a basis), then

yS
[r. )[
2
|r|
2
for all r H.
Equality holds for every r if and only if o is a basis.
10. GEOMETRY OF A HILBERT SPACE AND GRAM-SCHMIDT PROCESS 10-3
If the Hilbert space H is separable it contains a countable dense set then any
orthonormal basis is countable. In this case we can avoid Zorns Lemma. The foundation of
this is the Gram-Schmidt process.
Theorem 10.6 (Gram-Schmidt Process). Let
j
be a sequence of vectors in a Hilbert
space. Then there is an orthonormal sequence r
j
such that
span
1
. . . . .
n
spanr
1
. . . . . r
n
.
Proof. The proof is constructive. We may assume, without loss, that
1
. . . . .
n
is
linearly independent for each :. (Otherwise throw out vectors
n
until this is the case.)
Then
1
,= 0 so let
r
1
=

1
|
1
|
.
Clearly spanr
1
= span
1
.
Now, suppose we are given r
1
. . . . . r
n1
such that
spanr
1
. . . . . r
n1
= span
1
. . . . .
n1
.
Let
r
n
=

n1
j=1

n
. r
n1
) r
n1
_
_
_
n

n1
j=1

n
. r
n1
) r
n1
_
_
_
.
This is OK since
n
,=

n
j=1

n
. r
n1
) span
1
. . . . .
n1
. Clearly
spanr
1
. . . . . r
n
= span
1
. . . . .
n
.
By induction, the result follows.
Corollary 10.7. Let H be a separable Hilbert space. Then H has a countable orthonor-
mal basis.
Finally, let us discuss the isometries of Hilbert spaces.
Theorem 10.8. Let H and H
t
be Hilbert spaces. Given an orthonormal basis o for H,
an orthonormal set o
t
H
t
and a one-to-one onto map ) : o o
t
, dene a linear map
H H
t
via

yS

T
f

yS

y
)().
Then 1 is a linear isometry onto span o
t
H
t
. Furthermore, any isometry of H with a
subspace of H
t
is of this form.
Corollary 10.9. Two Hilbert spaces are isomorphic i their orthonormal bases have
equal cardinality. In particular, every Hilbert space is isomorphic with /
2
(o) for some set o.
Any separable, innite dimensional Hilbert space is isomorphic to /
2
.
Remark. For an arbitrary set o, /
2
(o) is dened to be the set of functions ) : o R
or C such that

yS
[)()[
2
< .
Note that ) /
2
(o) = : )() ,= 0 is countable.
The proof of these results is left as an exercise.
Part 4
Locally Convex Spaces
LECTURE 11
Locally Convex Spaces and Spaces of Test Functions
Reading: 13.-13.2 and B.1 of Lax.
A Banach space is one example of a topological vector space (TVS), which is a linear
space A together with a topology on A such that the basic operations of addition and scalar
multiplication are continuous functions.
Definition 11.1. A topological vector space is a linear space A with a Hausdor topol-
ogy such that
(1) (r. ) r + is a continuous map from A A (with the product topology) into
A.
(2) (/. r) /r is a continuous map from 1 A (with the product topolgoy, 1 = R or
C) into A.
Remark. Recall that a Hausdor space is one in which points may be separated by open
sets: given r. A, r ,= there are disjoint open sets l. \ , l \ = such that r l
and \ .
Theorem 11.1. Let A be a TVS and let l A be open. Then
(1) For any r A, l r = : +r l is open.
(2) For any scalar / ,= 0, /l = : /
1
l is open
(3) Every point of l is interior: given r l and A there is c 0 such that for
any scalar t with [t[ < c we have r +t l.
Proof. The set l r is the inverse image of l under the map + r. Thus
(1) follows from continuity of the map + r which follows from joint continuity of
(. r) +r. (Why?)
Likewise (2) follows from continuity of /
1
.
Since l r is open it suces to suppose r = 0 l. For xed A the map t t
is continuous. (Why?) Thus t : t l is open. Since this set contains t = 0 it must
contain an interval (c. c) (or an open ball at the origin if the eld of scalars is C).
The class of TVSs is rather large. However, almost all of the TVSs important to analysis
have the following property:
Definition 11.2. A locally convex space (LCS) is a TVS A such that every open set
containing the origin contains an open convex set containing the origin. That is, there is a
basis at the origin consisting of open convex sets.
Given a LCS, A, we dene the dual A
t
to be the set of all continuous linear functional
on A. A LCS space shares the property of separation of points by linear functionals:
Theorem 11.2. Let A be a LCS and let ,=
t
be points of A. There is a linear
functional / A
t
such that
/() ,= /(
t
).
11-1
11-2 11. LOCALLY CONVEX SPACES AND SPACES OF TEST FUNCTIONS
Proof. Of course, we use the Hahn-Banach theorem. Specically the hyperplane sepa-
ration Theorem 2.1.
First, it suces to suppose the eld of scalars is R, for if we construct a suitable real
linear functional /
r
on a complex LCS we can complexify it
/(r) = /
r
(r) i/
r
(ir).
Now, without loss we suppose that
t
= 0. Since the topology on A is Hausdor, there is
an open set l
t
with , l. Since A is locally convex, we may suppose l to be convex.
Replacing l with l (l) we may assume l is symmetric about 0, so r l = r l.
Since all points of l are interior, Theorem 2.1 asserts the existence of a linear functional /
with 1 = /() and /(r) < 1 for r l. In fact, the proof shows that
/(r) j
U
(r) r A.
where j
U
is the gauge function of l,
j
U
(r) = inft 0 : t
1
r l.
We need to show that / is continuous. It suces to show /
1
(c. /) is open for any
c < / R. Let t (c. /). Let r
0
be any point with /(r
0
) = t. Then, /
1
(c t. / t) =
/
1
(c. /) r
0
0. (Why?) Thus it suces to suppose c < 0 < / and show that /
1
(c. /)
contains an open neighborhood at 0. Let t = minc. /. The given r l,
/(tr) j
U
(tr) = tj
U
(r) < t and /(tr) j
U
(tr) = tj
U
(r) < t.
Thus tl /
1
(t. t) (c. /).
Converse to this construction is the following idea
Theorem 11.3. Let A be linear space and let 1 be any collection of linear functionals
on A that separates points: for any .
t
A there is / 1 such that /(r) ,= /(r
t
). Endow
A with the weakest topology such that all elements of 1 are continuous. Then A is a LCS,
and the dual of A is
A
t
= span 1 = nite linear combinations of elements of 1.
Remark. Recall that a topology on A is a collection | of subsets of A that includes
A and , and is closed under unions and nite intersections. The intersection of a family of
topologies is also a topology. Thus the weakest topology with property is the intersection
of all topologies with property .
Proof. Exercise.
What are some examples of LCSs? First o, any Banach space is locally convex, since
the open balls at the origin are a basis of convex sets. But not every LCS has a norm which
is compatible with the topology. By far the most important examples, though are so-called
spaces of text functions and their duals, the so-called spaces of distributions.
LECTURE 12
Generation of a LCS by seminorms and Frechet Spaces
Reading: 5.2 of Reed and Simon (or just these notes)
Test functions
The theory of distributions due to Laurent Schwarz is based on introducing a
LCS of test functions A and its dual A
t
, a space of distributions. The test functions
are nice: we can operate on them arbitrarily with all the various operators of analysis
dierentiation, integration etc. Using integration, we embed A A
t
via a map /

:
/

() =
_
R
d
(r)(r)dr.
Thus we think of a (test) function both as a map and as an averaging procedure. A key
identity is integration by parts
/

(
i
) = /

().
This suggests that we dene

i
/() = /()
for any distribution / A
t
. One typically uses a function notation for a distribution, writing
/() =
_
1(r)(r).
even if the function 1 doesnt exist.
A common, and useful, space of test functions is
C

c
(R
d
) = C

functions on R
d
with compact support..
We wish to topologize this set so that a sequence n
k
converges n if the supports of all n
k
are contained in some xed compact 1 R
d
and if for any choice of mult-index =
(
1
. . . . .
d
) N
d
we have
1

n
k
=

1
1

d
d
n
k
1

n uniformly in 1.
What is really going here is this. Dene for each : 0 and each multi-index a
semi-norm of C

c
(R
d
),
j
n,
(n) = sup
[x[2
n
[1

n(r)[ .
(Recall that a semi-norm on a linear space A is a map j : A [0. ) which is positive
homogeneous (j(cr) = [c[j(r)) and sub-additive (j(r + ) j(r) + j()). It is allowed
that j(r) = 0 for r ,= 0. ) It may happen that |n|
n,
vanishes even if n ,= 0, however the
collection separates points (Why?):
Definition 12.1. A collection of semi-norms o separates points if
j(n) = 0 j ( = n = 0.
12-1
12-2 12. GENERATION OF A LCS BY SEMINORMS AND FR

ECHET SPACES
Now endow A = C

0
(R
d
) with the smallest topology such that A is a TVS and each of
the semi-norms j
n,
is continuous.
Claim. C

c
(R
d
) with this topology is a LCS
Warning: this is not the standard topology on C

c
(R
d
). See below.
Generation of an LCS by semi-norms
The claim follows from the following general result.
Definition 12.2. Let A be a linear space and o a family of functions ) : A `, ` a
topological space (usually ` = R or C). The TVS topology generated by o is the weakest
topology on A such that A is a TVS and all the functions in o are continuous.
Theorem 12.1. Given a linear space A and a collection of semi-norms o that separates
points, the TVS topology generated by o makes A is locally convex.
Conversely, given a LCS A and ( a neighborhood base at the origin consisting of convex,
symmetric sets, the LCS topology on A is the TVS topology generated by o = j
U
: l (
are continuous.
Proof. First given an LCS space A let us show that the gauge function j
U
of a convex,
symmetric neighborhood of the origin l is continuous. To begin, note that
j
1
U
[0. /) = r A : r /l = /l
is open for each /. Next consider the sets j
1
U
(/. ). Let r be in this set and let = j
U
(r).
Consider the open neighborhood \ = r + ( /)l then for = r + ( /)
t
\ we have
j
U
() j
U
(r) ( /)j
U
(
t
) /.
So \ j
1
U
(/. ) and thus the set is open. Continuity of j
U
follows since the sets [0. /),
(/. ) as / ranges over (0. ) generate the topology on [0. ).
Now, any topology under which every j o is continuous certainly contains the collection
( = j
1
[0. /) : / (0. ).
consisting of convex, symmetric sets with the origin as an interior point. Consider the
smallest TVS topology containing this collection. It is easily seen to be locally convex. (We
need the fact that o separates points to get the Hausdor property.) Since j = j
U
for
l = j
1
[0. 1) we see from the above argument that all j o are continuous in this topology.
Thus this is the TVS topology generated by o.
Conversely, let T denote a given LCS topology on A. Thus T is certainly a topology un-
der which A is a TVS and all elements of o = j
U
: l a convex, symmetric neighborhood of 0
are continuous. To prove it is the weakest such, we must show that any such topology con-
tains T . Any l ( may be written as l = j
1
U
([0. 1)). Thus any topology under which
A is a TVS and all j
U
are continuous certainly contains ( and all its translates, and thus
T .
METRIZABLE LCSS 12-3
Metrizable LCSs
Above we claimed that the topology on C

0
could be given in terms of uniform conver-
gence of sequences of functions. However, in a general LCS sequential convergence may not
specify the topology a set may fail to be closed even it contains the limits of all conver-
gent sequences of its elements because there may not be a countable neighborhood base
at the origin. (Dont worry too much about this.) However, if the origin has a countable
neighborhood base then it turns out that the LCS is actually metrizable, so in particular
sequential convergence species the topology.
Theorem 12.2. Let A be an LCS. The following are equivalent
(1) A is metrizable
(2) A has a countable neighborhood basis at the origin ( consisting of convex, symmetric
sets
(3) the topology on A is generated by a countable family of semi-norms.
Proof. The equivalence of (2) and (3) is established by associating to convex, symmetric
neighborhoods of the origin the corresponding gauge function and vice versa. The details
are left as an exercise.
To show (1) = (2), suppose A is metrizable. Then A has a countable neighborhood
basis (this is a property of metric spaces), and in particular a countable neighborhood basis
at the origin. Since A is a LCS we may nd a convex, symmetric open set contained in each
of the basis sets, thus obtaining a convex, symmetric, countable neighborhood basis at the
origin.
To show (2) = (1), suppose A has a countable neighborhood basis as indicated, and
let T denote its topology. Since ( is countable we may assume, without loss, that it is
a decreasing sequence ( = l
1
l
2
. (Order the elements of ( and take nite
intersections l
k
l
1
l
k
.) Let j
j
(r) = j
U
j
(r), so j
k
(r) j
j
(r) if / ,. Dene a
metric
d(r. ) =

j=1
2
j
j
j
(r )
1 +j
j
(r )
.
and the metric topology T
d
. (Recognize this? Why is this a metric? Note that the collection
j
U
: l ( separates points since ( is a basis.)
Clearly,
r : d(r. 0) < 2
j1
l
j
.
Thus T T
d
(since any T open set contains a translate of some d-ball centered at each of
its points). On the other hand, if r tl
k
d(r. 0) <
k

j=1
2
j
t
1 +t
+

j=k+1
2
j

1
2
t
1 +t
+ 2
k1
.
Thus
2
k
l
k
r : d(r. 0) < 2
k
.
which shows that T
d
T . Thus T = T
d
and A is metrizable.
Definition 12.3. A Frechet space is a complete, metrizable, locally convex linear space.
Remark. Recall that a metric space is complete if every Cauchy sequence converges.
LECTURE 13
The dual of an LCS
It turns out that C

c
(R
d
), although a metric space with the metric suggested above,
is not complete. The completion is the space of C

0
functions that together with all their
derivatives vanish at , which could be topologized with the seminorms
j

()) = sup
x
[1

)(r)[ .
The correct way to think of C

c
(R
d
) is as an inductive limit, which is as the union

n
C

([2
n
. 2
n
]). There is an inductive limit topology on this space, making it a com-
plete, but non-metrizable, LCS. Each of the spaces C

([2
n
. 2
n
]) is a Frechet space, but
clearly the inductive limit topology is dierent that the (metrizable) LCS topology generated
above. It is still true that a sequence converges according to the criteria given above, but
this doesnt give the whole picture as the space isnt separable! (See Reed and Simon chapter
V.)
A remedy for this, that is often sucient, is to work with the following somewhat larger
space
Definition 13.1. The Schwarz space o(R
d
) consists of every C

function ) on R
d
such
that
j
,
()) = sup
xR
d
[r

)(r)[ < .
for every pair of multi-indices . N
d
.
Remark. r

= r

1
1
r

d
d
.
Note that o(R
d
) 1
p
(R
d
) for every j.
Likewise ) o(R
d
) = 1

) 1
p
(R
d
) for every j.
Corollary 13.1. The space o(R
d
) with the topology generated by the seminorms j
,
is a Frechet space.
The dual space of a LCS
It is useful to use the inner product notation to denote the pairing between elements of
A and linear functionals
/(r) = r. /) .
(Note that this inner product is linear in both factors even if we are dealing with complex
spaces.)
Given a linear space A and a linear space of linear functionals 1 on A that separates
points we have seen that there is a LCS topology on A such that 1 is the dual of A. This
topology is called the 1-weak topology on A and is denoted (A. 1).
13-1
13-2 13. THE DUAL OF AN LCS
On the other hand, given an LCS, we can think of A as a collection of linear functionals
on A

, associating to r A the map


/ r. /) .
The A-weak topology on A

, (A

. A), is also called the weak

toplogy. It is generated by
the family of seminorms
j
x
(/) = [r. /)[ .
Theorem 13.2. If A is an LCS then (A

. (A

. A))

= A.
Remark. Recall that (A. A

) is the given LCS topology on A so we also have (A(A. A

))

=
A

. Thus for any LCS (A

= A, provided we topologize A

with the weak

toplogy. If A
is a Banach space we also have a norm topology on A

, which is substantially stronger than


the weak

topology and with respect to which this identity may not hold. For instance,
(1) As Banach spaces c

0
= /
1
and /

1
= /

and /

, which includes Banach limits, is


strictly larger than /
1
.
(2) As LCS spaces c

0
= /
1
and (/
1
. (/
1
. c
0
)

= c
0
, etc.
The moral of the story is topology matters.
The following theorem is useful for determining if a linear functional is continuous.
Theorem 13.3. Let A be a LCS generated by a family of semi-norms o. Then a linear
functional / A
t
if and only if there is a constant C 0 and a nite collection j
1
. . . . . j
n
o
such that
[/(r)[ C
n

j=1
j
j
(r) r A.
Proof. () If / is continuous then l = /
1
(1. 1) an open, convex, symmetric neigh-
borhood of the origin in A. By virtue of the fact that o generates the topology on A, since
l is open we have
n

j=1
r : j
j
(r) < l
for some nite collection j
1
. . . . . j
n
. Thus,
\ = r :
n

j=1
j
j
(r) < l.
Now, given r let t =
2

n
j=1
j
j
(r). Then
n

j=1
j
j
(t
1
r) =

2
.
so t
1
r \ l. Thus

/(t
1
r)

< 1 =

n
j=1
j
j
(r)

n
j=1
j
j
(r)
=
2

j=1
j
j
(t
1
r).
Multiplying through by t we get the desired bound with C = 2,.
THE DUAL SPACE OF A LCS 13-3
() Since / is linear, it suces to show that / is continuous at 0. That is we must show
that /
1
(. ) contains an open set containing the origin for each 0. But clearly
r : C
n

j=1
j
j
(r) < /
1
(. ).

LECTURE 14
Spaces of distributions
Tempered Distributions
The dual of o(R
d
) denoted o

(R
d
) is the space of tempered distributions. Here are some
examples:
(1) o(R
d
) o

(R
d
) where we associate to a function o(R
d
) the distribution
. ) =
_
R
d
(r)(r)dr.
(2) More generally, a function 1 1
1
loc
(R
d
) that is polynomially bounded in the sense
that j(r)
1
1(r) 1
1
(R
d
) for some positive polynomial j 0 may be considered as
a tempered distribution
j:i. 1) =
_
R
d
(r)1(r)dr.
(3) Similarly, any polynomially bounded Borel measure j, with
_
j(r)
1
d[j[(r) < .
is a tempered distribution:
. j) =
_
R
d
(r)dj(r).
To go further we need the following generalization of Theorem 13.3 from the last lecture:
Theorem 14.1. Let A. 1 be a LCSs generated by a families of semi-norms o. T respec-
tively. Then a linear map 1 : A 1 is continuous if and only if for any semi-norm o
there is a constant C 0 and a nite collection j
1
. . . . . j
n
o such that
(1r) C
n

j=1
j
j
(r) r A.
Corollary 14.2. For each , = 1. . . . . d, dierentiation
j
is a continuous map from
o o.
Now we dene
j
: o

. Note that

j
. 1) = .
j
1) .
whenever 1 is a C
1
function of polynomial growth. Thus dene for arbitrary / o

:
.
j
/) =
j
. /) .
Proposition 14.3. So dened,
j
: o

is a continuous map.
Proof. Exercise
14-1
14-2 14. SPACES OF DISTRIBUTIONS
Thus we have the following generalization of the above examples:
Let ^
d
be a multi-index and let 1 be a polynomial 1
1
bounded function. Then
1

1 is a tempered distribution:
. 1

1) = (1)

. 1) = = (1)

_
R
d
1

(r)1(r)dr.
Theorem 14.4 (Structure Theorem for Tempered Distributions). Let / o(R
d
) be a
tempered distribution. Then there is a polynomially bounded continuous function p and a
multi-index N
d
such that / = 1

p.
For the proof see Reed and Simon, Ch. V.
For example, we now understand in a precise sense the identity
(r) =
d
2
dr
2
[r[.
More generally, in d = 2, we have the identity
(r) =
1
2
ln [r[.
This can be veried as follows. Note that ln [r[ = 0 if r ,= 0. Thus
. ln [r[) = 0 if (r) = 0 for [r[ < c.
Let / be a compactly supported function that is 1 in the neighborhood of the origin, then
/. ln [r[) = 0. Thus it suces to suppose C

c
(R
d
). For such
. ln [r[) =
_
[x[<M
(r) ln [r[dr = lim
0
_
<[x[<M
(r) ln [r[dr.
since ln [r[ is locally integrable. Now integrate by parts:
_
<[x[<M
(r) ln [r[dr =
_
<[x[<M
r
[r[
2
dr ln
_
[x[=
r
[r[
(r)d(r).
where in the rst integral
x
[x[
2
is ln [r[ and in the second integral
x
[x[
is an outward facing
normal on [r[ = and d(r) is the length measure on the circle. Continuing, we nd that
_
<[x[<M
(r) ln [r[dr = =
1

_
[x[=
(r)d(r) C( ln ) (0).
Remark. In a similar fashion, in dimension d 3,
(r) = c
d
[r[
2d
.
with c
1
d
= (d 2) area of the unit sphere [r[ = 1 : r R
d
.
Other spaces of distributions
Consider the scale of spaces
C

c
(R
d
) o(R
d
) C

0
(R
d
) C

(R
d
).
Each has a natural LCS topology such that it is a complete space. The middle two are
Frechet spaces and we have discussed their topologies already.
OTHER SPACES OF DISTRIBUTIONS 14-3
On C

c
(R
d
) we put the inductive limit topology as follows. For each :, let
n
=
(2
n
. 2
n
)
d
. The space A
n
= C

0
(
n
) is a Frechet space generated by the seminorms
j
n,
(n) = sup
xn
[1

n(r)[ . (-)
Note that A
n
A
n+1
, the embedding is continuous, and A
n
is a closed subspace of A
n+1
.
Put of C

c
(R
d
) = A =
n
A
n
the LCS topology which has the following convex neighborhood
base at the origin
( = l A : l is convex, symmetric, all points of l are interior, and lA
n
is open for each :.
Theorem 14.5. With this topology C

c
(R
d
) is a LCS and is complete. The topology does
not depend on the choice of sets
n
and a sequence n
n
C

c
(R
d
) converges if and only if
there is ` such that n
n
A
N
for all : and n
n
converges in A
N
.
For the proof see Reed and Simon.
A distribution is an element of T

(R
d
) = (C

c
(R
d
))

. Since any tempered distribution 1


acts on C

c
(R
d
) and furthermore if . 1) = 0 for all in C

c
then 1 = 0 (note that C

c
(R
d
)
is dense in o), we have the embedding o

T. Distributions in T need not be bounded at


any locally integrable function, like e
e
e
|x|
2
is a distribution. Also, the structure theorem
doesnt hold in this context, however we have
Theorem 14.6. Let 1 be a distribution in T(R
d
) then there is a sequence )

of continuous
functions, indexed by multi-indices, such that
1 =

.
where for any n C

c
(R
d
) only nitely many terms contribute to the sum
n. 1) =

n. 1

) .
Roughly speaking, the order of the distribution can become unbounded at .
Likewise we may put a topology on C

(R
d
) given by uniform convergence on compact
subsets. That is generated by the family of semi-norms (-), where now n need not have
compact support. This makes c(R
d
) = C

(R
d
) into a Frechet space. Let us denote its dual
by c

. We have the inclusions


c

0
o

.
where T

0
is the dual of C

0
(R
d
). Distributions in c

have compact support, where


Definition 14.1. A distribution vanishes on an open set l R
d
if . 1) = 0 whenever
has compact support in l. The support of a distribution is the smallest closed set 1 such
that 1 vanishes on R
d
1.
Distributions in T
0
are bounded, in the sense that they may be written as nite
sums of derivatives of nite measures. Distributions in o are unbounded at , but only
polynomially so.
LECTURE 15
Applications: solving some PDEs
The Poisson equation
The electric potential produced by a charge distribution is known to satisfy the
Poisson equation:
(r) = (r). r R
3
with the permitivity of space set = 1. Let us use the fact that

1
4
[r[
1
= (r)
to solve this equation. We need the following notion.
Definition 15.1. Given two functions . the convolution of and is
(r) =
_
R
d
(r )()d.
whenever the integral is dened.
Proposition 15.1. The convolution product is abelian, = and is a continuous
map of 1
1
1
1
1
1
, o o o. and T T T.
Proof. Exercise.
Theorem 15.2 (Youngs Inequality). Convolution is a continuous map from 1
p
1
q

1
r
with 1,j + 1, = 1 + 1,:, 1 j. . : .
Proof. Let ) 1
r

, 1,:
t
+ 1,: = 1. Then
[
_
)(r) (r)dr[
_ _
[)(r)(r )()[drd.
Write the integrand as (r. )(r. )(r. ) with
(r. ) = [)(r)[
r

[(r )[
q
p

(r. ) = [(r )[
q
r
[()[
p
r
(r. ) = [)(r)[
r

[()[
p
q

and 1,j + 1,j


t
= 1, 1, + 1,
t
= 1. Note that 1,j
t
+ 1,
t
= 1,:
t
= 1 1,:. So, by Holder,
_ _
[)(r)(r )()[drd
__ _
[)(r)[
r

[(r )[
q
drd
_ 1
p

__ _
[(r )[
q
[()[
p
drd
_1
r
__ _
[)(r)[
r

[()[
p
drd
_ 1
q

= |)|
r

+
r

||
q
p

+
q
r
q
||
p
r
+
p
q

p
.
15-1
15-2 15. APPLICATIONS: SOLVING SOME PDES

Now note that if C

and ) is integrable and compactly supported, say, then )


is C

. Indeed

j
( )) = (
j
) ).
Likewise, if o and ) is integrable with compact support or more generally polynomial
decay at , so j(r)) 1
1
for any polynomial j then ) o. If also o then
. )) =
_


).
_
.
where

)(r) = )(r).
Thus we dene the convolution 1 ) of a ), a compactly supported integrable function, and
a distribution 1 to be the distribution
. 1 )) =
_


). 1
_
.
Theorem 15.3. If ) 1
1
has compact support and 1 is a tempered distribution, then
1 ) o

is a tempered distribution, and if 1 1


1
loc
o

then 1 ) 1
1
loc
as well and
. 1 )) =
_
R
d
R
d
(r)1(r ))()drd.
The map . 1 1 is a continuous map of o o

.
Proof. The rst two statements are clear. To show that 1 is C

if o, note that
this follows if 1 is a function since

j
_
(r )1()d =
_

j
(r )1().
More generally, we may write a general distribution as 1

p with p continuous, and integrate


by parts to see that
1 (r) =
_
1

(r )p()d.
Continuity of the map is easy.
Now, let 1(r) =
1
4
[r[
1
. This is a locally integrable function, so if 1
1
loc
we have
(r) =
1
4
_
1
[r [
()d
as a (weak) solution of the Poisson equation. It is not the unique solution since we can add
to it any harmonic function l which satises l = 0. For example, l(r) = r
1
is harmonic.
So is l(r) = r
1
r
2
. In fact, one can show that any Harmonic tempered distribution is a
polynomial. (The easy way to do this is to use the Fourier Transform.)
More generally we have the following theorem due to Weyl
Theorem 15.4 (Weyl). Let 1 be a distribution that satises 1 = 0 on an open set
1 R
n
, so . 1) = 0 for C

c
(1). Then 1 is a C

function in 1: there is
p C

(1) such that . 1) = . p) for C

c
(1).
For the proof, see appendix B.
LECTURE 16
The Dirichlet problem
Reading: 7.2 of Lax
Suppose we want to solve Poissons equation for that vanishes outside an open set
1 R
d
containing the support of . That is
_
(r) = (r) r 1
(r) = 0 r , 1.
It is useful to use Hilbert space methods. Let us set up the inner products
). p)
0
=
_
D
)(r)p(r)dr and ). p)
1
=
_
D
d

j=1

j
)(r)
j
p(r)dr.
Take these over C

c
(1), which is incomplete but completes to 1
2
(1) with respect to the
rst and to a Sobolev space in the second.
Lemma 16.1. For any ) C

c
(1)
|)|
0

diam(1)

d
|)|
1
where diam(1) = sup
x,yD
[r [ is the width of 1.
Proof. This is a calculus exercise. Extend ) to be identically 0 outside 1. Note that
)(r) =
_
x
1
x
(0)
1

1
)(. r
2
. . . . . r
d
)d.
where r
(0)
= (r
(0)
1
. r
2
. . . . . r
d
) 1. Thus by Cauchy-Schwarz,
)(r)
2
diam(1)
_
R

1
)(. r
2
. . . . . r
d
)d.
Integrating over r 1 now gives
|)|
2
0
diam(1)
2
|
1
)|
2
0
.
Similarly,
|)|
2
0
diam(1)
2
|
j
)|
2
0
.
Averaging these results gives the Lemma.
Let H
(0)
1
denote the completion of C

0
(1) in the norm ||
1
. This is a Hilbert space.
Lemma 16.2. Every element ) H
(0)
1
may be identied with a locally integrable function
such that
(1) )(r) = 0 if r , 1
(2) ) 1
2
(1)
16-1
16-2 16. THE DIRICHLET PROBLEM
(3) For , = 1. . . . . d,
j
) (in the sense of distributions) is a locally integrable function
in 1
2
(1).
Furthermore we have
). p)
1
=
D

j=1

j
).
j
p)
0
and
).
j
p)
0
=
j
). p)
whenever ). p H
(0)
1
.
Proof. The inequality |)|
0
const.. |)|
1
shows that any Cauchy sequence of function
in the ||
1
norm is Cauchy in 1
2
. So any sequence )
n
C

c
(1) which converges in the H
(0)
1
also converges in 1
2
. Identify the limit ) with the corresponding element of 1
2
(1), extended
to be zero outside 1. (1) and (2) follow.
To derive (3), note that for p C

c
p.
j
)
n
) =
j
p. )
n
)
by integration by parts. The r.h.s. converges to
j
p. )) using Cauchy-Schwarz. On the other
hand,
j
)
n
converges in the 1
2
norm to a function /
j
. We conclude that /
j
=
j
) in the
sense of distributions, so
j
) 1
2
(1).
To show the formulas, note that they hold for elements of C

c
(1) and follow for the
). p H
(0)
1
by taking limits.
Now, x 1
2
(1). This function gives rise to a linear functional / on H
(0)
1
1
2
(1) by
/(n) = n. ))
0
.
Since
[/(n)[ |n|
0
|)|
0

diam(1)

d
|)|
0
|n|
1
.
this is a bounded linear functional. By Riesz-Frechet there is an element H
(0)
1
such that
n. )
0
= n. )
1
.
for all n H
(0)
1
. Specializing to n C

c
(1), we nd that
) =
in the sense of distributions.
In a similar way, we may write down a bilinear form
1(n. ) =
_
D
_

i,j

i,j
(r)
i
n(r)
j
(r) +n(r)

i
1
i
(r)
i
(r) +:(r)n(r)(r)
_
dr.
Let us check the hypotheses of Lax-Milgram. Clearly 1 is bilinear and bounded. To estimate
1(n. n) from below, suppose that
i,j
(r) is pointwise positive denite, that is

i,j

i,j
(r)
i

j
(r)

2
i
.
Then
1(n. n)
_
D
_
(r) [n(r)[
2
[n(r)[ [F(r)[ [n(r)[ +:(r)n(r)
2
_
dr.
16. THE DIRICHLET PROBLEM 16-3
Now suppose the quadratic function
(r)t
2
[F(r)[ t: +:(r):
2
(t
2
+:
2
).
uniformly in r, for some 0. This amounts to the requirement that (r) 0, :(r) 0
and the discriminant
sup
x
[F(r)[
2
4(r):(r) < 0.
Then
1(n. n) (|n|
2
1
+|n|
2
0
) |n|
2
1
.
Thus, by Lax-Milgram given 1
2
(1) we may nd H
(0)
1
such that
. )
0
= 1(. ) H
(0)
1
.
Such is a distributional solution to the equation

i,j

i,j
(r)
j
(r) +

i
1
i
(r)
i
(r) +:(r)(r) = (r). r 1.
One might ask, in what sense do the elements of H
(0)
1
vanish outside 1? The complete
answer to this question involves the theory of Sobolev spaces. An elementary answer is the
following:
Lemma 16.3. Let 1 be a hypercube [1. 1]
d
and let ) H
(0)
1
. Then
lim
0
1
[dist(r. 1) < [
_
dist(x,D)<
[)(r)[ = 0.
Remark. [ [ denotes Lebesgue measure, so [dist(r. 1) < [ = C(),
Proof. The set of points within distance of 1 is a union of slabs such as o = r
1 : 1 < r
1
< 1+. Let ) C

c
(1) and consider the integral of [)(r)[ over this slab.
Integrating by parts, we have
_
S
[)(r)[dr =
_
S

1
[)(r)[(1 + r
1
)dr.
since ) vanishes on the boundary. Now
1
[)(r)[ [
1
)(r)[. Thus,
_
S
[)(r)[dr
_
S
[
1
)(r)[(1 + r
1
)dr [o[
1
2
|)|
1
C |)|
1

3
2
.
since [o[ = C(). Adding up the contributions from all slabs we get
_
dist(x,D)<
[)(r)[dr C |)|
1

3
2
.
Taking limits, this estimate follows, with constant C independent of ), for ) H
(0)
1
. Since
[dist(r. 1) < [ c the result follows.
Remark. We were a bit wasteful. We could have averaged over just

= dist(r. ) <
with a piece of the boundary with codimension 2. In that case [

[
2
and
_

[)(r)[dr C
2
__

[)(r)[
2
dr
_1
2
.
The integral on the r.h.s. goes to zero, so the average of [)[ vanishes also on codimension 2
sets.
Part 5
Weak Convergence and Weak Topology
LECTURE 17
Dual of a Banach space
Reading: 8.2 -8.3 of Lax.
Recall that the dual of a Banach space A is the space A
t
of bounded linear functionals
on A. Recall that we put a norm topology on A
t
by dening
|/| = sup r ,= 0
[/(r)[
|r|
.
Likewise, we have the following dual characterization of the norm on A:
Theorem 17.1. For every r A we have
|r| = max / ,= 0. / A
t
[/(r)[
|/|
.
Proof. Since |/| || [/()[ the l.h.s. is no smaller than the r.h.s. Thus we need only
produce an / such that [/(r)[ = |r| |/|. Dene / rst on the one dimensional subspace
spanr by /(tr) = t |r|. Since this functional is norm bounded by 1 on this subspace it
has an extension (by Hahn-Banach) to the whole space with this property.
We also have the weak

topology on A
t
, the weakest LCS topology such that all elements
of A are continuous functionals on A
t
. Recall that (A
t
. wk

= A. One might wonder if


(A
t
)
t
= A as Banach spaces. In fact this does not hold in general.
Definition 17.1. A Banach space is called reexive if (A
t
)
t
= A. That is if every
bounded linear functional on A
t
is of the form / /(r) for some r A.
Many important spaces are reexive, but not all. For instance:
Theorem 17.2. c
t
0
= /
1
, /
t
1
= /

, and /
t

/
1
.
Proof. Let / be a linear functional on c
0
. Evaluating / on the sequence e
k
(:) = 1 if
/ = : and 0 otherwise produces a sequence
b(:) = /(e
n
).
If a c
0
then
a = lim
n
n

j=1
c(,)e
j
.
so by linearity and continuity
/(a) = lim
n
n

j=1
c(,)b(,).
17-1
17-2 17. DUAL OF A BANACH SPACE
That is for any a C
0
the sequence ab is summable and
/(a) =

j=1
a(,)b(,).
To see that b /
1
, take a
n
(,) = e
i arg b(j)
if , : and 0 otherwise. Thus
n

j=1
[b(,)[ = /(a
n
) |/| .
It follows that b /
1
. (Conversely, it is clear that any b /
1
gives a linear functional on
c
0
.)
The same idea works to prove /
t
1
= /

. Finally, it is clear that /


1
/
t

, however no
Banach limit such as dened in the third lecture can be written as a scalar product with
something in /
1
.
However,
Theorem 17.3. Every Hilbert space is reexive.
Proof. In this case A
t
= A by Riesz-Frechet.
For 1
p
spaces we have.
Theorem 17.4. Let A. j be a nite measure space. For 1 j < the dual of 1
p
(A)
is 1
q
(A) where
1
p
+
1
q
= 1.
Remark. The result, but not the proof, extends to arbitrary measure spaces for 1 <
j < . If j = 1 (so = ) the result holds in -nite measure spaces.
Proof. The Holder inequality shows that 1
q
(A) 1
p
(A)
t
, via the pairing
). p) =
_
X
)(r)p(r)dj(r). ) 1
p
and p : 1
q
.
Suppose now that / is a linear functional on 1
p
(A). For each measurable set we have

A
1
p
(A). So
() = /(
A
)
denes a nitely additive set function. In fact, it is countably additive since / is continuous,
and if
1
.
2
. . . . are pairwise disjoint then
n

j=1

A
j

A
j
in 1
p
as one may readily show. Furthermore () = 0 if j() = 0 since then
A
= 0 in 1
p
. Thus
<< j and by Radon-Nikodym there is function p 1
1
such that
/(
A
) = () =
_
A
p(r)dj(r) =
_
X

A
(r)p(r)dj(r).
By taking limits of simple functions we have
/()) =
_
A
)(r)p(r)dj(r) (-)
whenever ) 1

, using dominated convergence.


17. DUAL OF A BANACH SPACE 17-3
It remains to show that p 1
q
for then (-) extends to all of 1
p
by density of 1

. To
this end, x t 0 and let
)
t
(r) =
_
[g(x)[
q
g(x)
0 < [p(r)[ < t
0 [p(r)[ = 0 or [p(r)[ t.
Clearly [)
t
(r)[ t
q1
so )
t
1

1
p
. Thus
_
[g(x)[<t
[p(r)[
q
dj(r) =
_
)
t
(r)p(r)dj(r) |/|
__
[)
t
(r)[
p
dj(r)
_1
p
.
But [)
t
(r)[
p
= [p(r)[
pqp
= [p(r)[
q
on [p(r)[ < t. Thus
__
[g(x)[<t
[p(r)[
q
dj(r)
_
1
1
p
|/|
and the result follows.
Corollary 17.5. 1
p
(A) is reexive for 1 < j < .
This result also follows from
Theorem 17.6 (Milman 1938). Any uniformly convex Banach space is reexive
In general 1
1
is not reexive: (1
1
)
t
= 1

but 1

contains linear functionals that are


not in 1
1
. Notice where the proof breaks down. Given a linear functional / on 1

we can
dene a set function
() = /(
A
)
as above. It is certainly additive, and absolutely continuous since if j() = 0 then
A
= 0
in 1

. It is not, however, countably additive since


n

j=1

A
j
,
A
j
in 1

.
The inequality (1

) ,= 1
1
also follows from:
Theorem 17.7. Let A be a Banach space over C. If A
t
is separable so is A.
Proof. Let /
n
be a countable dense subset of A
t
. For each : there is r
n
A such
that
|r
n
| = 1 and /
n
(r
n
)
1
2
|/
n
| .
It suces to show spanr
n
is dense in A.
Suppose contrarily that spanr
n
, = A. Then there is a non-zero linear functional / A
t
such that /(r
n
) = 0 for all :. We may assume that |/| = 1. However, we can nd : such
that |/ /
n
|
1
4
, say. Thus |/
n
|
3
4
and
0 = /(r
n
) = /(r
n
) /
n
(r
n
) +/
n
(r
n
)
1
2
|/
n
| |/ /
n
|
1
8
.
Thus no such / exists and we must have spanr
n
= A.
LECTURE 18
Riesz-Kakutani theorem
Theorem 18.1 (Riesz-Kakutani). Let Q be a compact Hausdor space, C(Q) the space
of continuous real valued functions on Q with the max norm. Then C(Q)
t
= `(Q) = set of
signed Borel measures of nite total variation.
That is to every bounded linear functional / C(Q)
t
is associated a unique Borel measure
: such that
/()) =
_
Q
)d:.
Furthermore the norm of / is the total variation |/| =
_
Q
[d:[.
Remark. i) A Borel measure is a measure on sets in the Borel -algebra which is the
smallest -algebra containing the open sets on Q. ii) The dual of C(Q; C) is the set of
complex Borel measures, : = :
r
+ i:
i
with :
r
. :
i
`(Q). iii) Compactness is crucial
here. (More on this later.)
Before proving this theorem, let us prove
Theorem 18.2. Given / C(Q) there is a unique deomposition / = /
+
/

with /

positive linear functionals and |/| = /


+
(1) +/

(1).
Proof. Let C(Q)
+
= set of non-negative functions in C(Q). For ) C(Q)
+
dene
/
+
()) = sup/(/) : 0 / ).
It is clear that /
+
(t)) = t/
+
()) for t 0 and that /
+
()) /(0) = 0 for ) C(Q)
+
. Given
), p in C(Q)
+
their sum is also in C(Q)
+
. Clearly
/
+
() +p) sup/(/
1
) +/(/
2
) : 0 /
1
) and 0 /
2
p = /
+
()) +/
+
(p).
The opposite inequality clearly follows from the following result
Claim. Given ). p C(Q) and 0 / ) +p we can write / = /
1
+/
2
with /
1,2
C(Q),
0 /
1
) and 0 /
2
p.
Proof of Claim. Let /
1
= min). /. So /
1
is continuous and 0 /
1
). Let
/
2
= / /
1
. Since /
1
/, /
2
0. When /
1
= ) we have /
2
) +p ) = p. On the other
hand if /
1
= / we have /
2
= 0 so /
2
p.
Thus /
+
(t) + :p) = t/
+
()) + :/
+
(p) whenever t. : 0 and ). p C(Q)
+
. Dene /
+
on
all of C(Q) by
/
+
()) = /
+
()
+
) /
+
()

). with )
+
= max). 0 and )

= min). 0.
It is not hard to see that /
+
is linear. Now set /

= /
+
/ and note that
/

()) = sup/(/) /()) : 0 / ) = sup/(/) : ) / 0 0.


for ) C(Q)
+
.
18-1
18-2 18. RIESZ-KAKUTANI THEOREM
To prove |/| = /
+
(1) +/

(1) note rst that |/

| = /

(1) since these are positive linear


functionals. Thus, |/| |/
+
| +|/

| = /
+
(1) +/

(1). On the other hand by the denition


of /
+
and the corresponding inequality for /

we have
/
+
(1) +/

(1) = sup/(/) : 0 / 1 + sup/(/) : 1 / 0


= sup/(p) : 1 p 1 |/| .
We are now ready to prove the Riesz-Kakutani Theorem. By the splitting of a linear
functional into positive and negative parts, it suces to show
Theorem 18.3 (Riesz-Kakutani). Let / be a positive linear functional on C(Q) with Q
a compact Hausdor space. Then there is a unique positive Borel measure : on Q such that
/()) =
_
Q
)(r)d:(r) and |/| = :(Q). Conversely, any positive Borel measure : gives a
positive linear functional on C(Q) via ). :) =
_
Q
)(r)d:(r).
Proof. First, it is clear that any positive Borel measure : gives rise to a positive linear
functional by /()) =
_
Q
)(r)d:(r). (Continuous functions are Borel measurable.) Since
the functional is positive |/| = /(1) = :(Q).
It remains to show that every positive linear functional is of this form. So let / C(Q)
t
be
given. We would like to dene :(o) = /(
S
). for o a measurable set. However,
S
, C(Q)
(unless o is both open and closed, so a union of connected components). Thus we do the
next best thing: given an open set l we take
:(l) = sup/()) : ) l.
where, ) l indicates that ) C(Q), supp ) l and 0 ) 1.
Claim. Let l
1
. l
2
Q be a disjoint pair of open sets. Then :(l
1
l
2
) = :(l
1
)+:(l
2
).
Proof of claim. Note that given )
j
l
j
, we have )
1
+ )
2
l
1
l
2
and conversely
given ) l
1
l
2
, we have )
U
j
l
j
. The identity follows.
Thus : is nitely additive on open sets. So far the construction works on an arbitrary
compact space, but the open sets do not form a -algebra. A rst step is to dene : on
closed sets 1 by
:(1) = :(Q) :(1) = inf/()) : 1 ).
where 1 ) if 0 ) 1 and )(r) = 1 on 1.
To proceed we need the following result which uses the assumption that Q is a compact
Hausdor space:
Claim. If 1 l, 1 closed and l open, then
:(1) = inf/()) : 1 ) l sup/()) : 1 ) l = :(l).
Proof of claim. It is clear that
:(1) inf/()) : 1 ) l sup/()) : 1 ) l :(l).
To obtain equality on the two ends, we use Urysohns lemma, valid here since a compact
Hausdor space is normal (disjoint closed sets may be separated by open sets):
Theorem 18.4 (Uryshohns Lemma). Given two disjoint closed sets 1
0
. 1
1
Q there is
a continuous function ) : Q [0. 1] such that ) 0 on 1
0
and ) 1 on 1
1
.
18. RIESZ-KAKUTANI THEOREM 18-3
Given )
1
l, let 1
1
= 1 supp )
1
, and let \ be open with 1
1
\ 1 l such
\ exists because Q is normal. Now let 1
0
= \
c
. Then ) , on 1
j
, , = 0. 1 clearly satises
)
1
) and 1 ) l. We conclude that
sup/()) : 1 ) l = sup/()) : ) l = :(l).
Similarly given )
1
~ 1, let 1
1
= 1 and 1
0
= l
c
. Pick )
2
C(Q) with 1 )
2
l and
set ) = )
1
)
2
. Clearly 1 ) l and ) )
1
. Thus /()) /()
1
) so
inf/()) : 1 ) l = inf/()) : 1 ).
It now follows that
Claim. If 1 is closed then :(1) = inf:(l) : 1 l and l open.
If l is open then :(l) = sup:(1) : 1 l and 1 closed.
Proof of claim. Clearly :(1) inf:(l) : 1 l. To prove the converse, note
that given c 0 there is ) ~ 1 such that /()) :(1) + c. Let l = r : )(r)
1 c. Note that p l = p
1
1
) so /(p)
1
1
/()). Thus :(l)
1
1
/()). Thus
(1 c):(l) :(1) +c. It follows that inf:(l) : 1 l :(1).
The opposite identity, :(l) = sup:(1), follows by taking complements.
Now given arbitrary o Q dene
:
+
(o) = inf:(l) : o l and l open.
:

(o) = sup:(1) : o 1 and 1 closed.


Clearly :

(o) :
+
(o). If these two numbers are equal dene
:(o) = :

(o) if :
+
(o) = :

(o).
Claim. The collection = o Q : :
+
(o) = :

(:) is a -algebra containing all


Borel sets and : denes a countably additive measure on this -algebra.
Proof of Claim. Clearly contains all closed and all open sets. Thus once we show
it is a -algebra it is immediate that it contains all Borell sets. Clearly o = o
c
.
Thus we need only show that is closed under countable unions. This is left as an exercise
as is countable additivity of :.
To complete the proof, we must show that
_
)d: = /()). It suces to prove this for )
with 0 ) 1. Note that
_
Q
)d: =
_
1
0
:r : )(r) tdt =
_
1
0
:r : )(r) tdt.
Let us show that
_
)d: /()). Note that
_
Q
)d:
n

j=1
1
:
:
_
r : )(r)
, 1
:
_
.
Let p
j;n
be functions with
_
)(r)
j1
n
_
p
j;n

_
)(r)
j2
n
_
. Such functions exist by
Urysohns Lemma. So
_
Q
)d:
n

j=1
1
:
/(p
j;n
) = /(
n

j=1
1
:
p
j;n
).
18-4 18. RIESZ-KAKUTANI THEOREM
Now for every r Q
n

j=1
1
:
p
j;n
(r) =
]nf(x)|+1

j=1
1
:
+C(
1
:
) =
1
:
:)(r)| +C(
1
:
).
We conclude that

j=1
1
:
p
j;n
(r) )(r)

= C(
1
:
).
so
/(
n

j=1
1
:
p
j;n
) /()).
and
_
)d: /()).
To show the reverse inequality, note that
_
Q
)d:
n

j=1
:
_
)(r)
,
:
_
.
Thus,
_
Q
)d: /(
n

j=1
1
:
/
j;n
).
with )(r)
j+1
n
/
j;n
)(r)
j
n
. Then
n

j=1
1
:
/
j;n
(r) =
]nf(x)|1

j=1
1
:
+C(
1
:
) =
1
:
:)(r)| +C(
1
:
).
Again

j
/
j;n
), so
_
)d: /()). completing the proof.
A word about the non-compact case. Suppose A is a locally compact Hausdor space,
so A is a Hausdor space such that every point is contained in an open set with compact
closure. Then we can consider several spaces C
c
(A) C
0
(A) C
b
(A) C(A) where
C
b
(A) = bounded continuous functions. The middle two are Banach spaces in the sup
norm. The rst and last are LCS spaces: C
c
(A) has an inductive limit topology obtained
from writing it as C
c
(A) =
U
C
0
(l) where the union is over open sets l with compact
closure, C(A) has a topology generated by the seminorms j
K
()) = sup
xK
[)(r)[ for compact
1. Regarding the duals of these spaces, we have
C
c
(A)
t
= `(A) = Borel measures : such that [:[(1) < for any compact 1.
C
0
(A)
t
= `
0
(A) = nite Borel measures on A.
C
b
(A)
t
= nite Borel measures on the Stone-

Cech compactication of A.
C(A)
t
= compactly supported Borel measures.
LECTURE 19
Weak convergence
Reading: 10.1 and 10.2 of Lax.
We have already dened weak topologies in the general context of LCS spaces. Let us
now look at them in the special case of Banach spaces. Consider a Banach space A and
its Banach space dual A
t
. So A
t
is a collection of linear functionals on A. The weakest
topology on A so that every element of A
t
is continuous is called the weak topology on A,
denoted (A. A
t
).
A sequence r
n
in A converges weakly to r converges in the weak topology if
/(r
n
) /(r) for every / A
t
.
Sometimes this is denoted
r
n
r.
or
wk lim
n
r
n
= r.
This notion is weaker than strong convergence, which is in norm:
|r
n
r| 0. or r
n
r.
That is more sequences converge weakly than converge strongly. For instance

[n,n+1]
0 in 1
p
(R). 1 < j < .
but
_
_

[n,n+1]
_
_
L
p
= 1.
(Note that
[n,n+1]
does not converge weakly in 1
1
.)
Proposition 19.1. Let r
n
be an orthonormal sequence in a Hilbert space H. Then
r
n
0.
Remark. Since |r
n
| = 1, r
n
does not converge strongly to 0.
Proof. Fix H. By Bessels inequality

n
[ . r
n
) [
2
||
2
.
we see that . r
n
) 0. By the Riesz theorem on linear functionals on a Hilbet space,
r
n
0.
Theorem 19.2. Suppose r
n
A, a Banach space, satises
(1) r
n
are uniformly bounded: sup
n
|r
n
| < .
(2) lim/(r
n
) = /(r) for / 1
t
with 1
t
dense in A
t
.
Then r
n
r.
Proof. This is an easy approximation argument and is left as an exercise.
19-1
19-2 19. WEAK CONVERGENCE
The interesting thing is that the converse is true: weakly convergent sequences are
uniformly bounded. To prove this we will use
Theorem 19.3 (Principle of Uniform Boundedness for a complete metric space). Let A
be a complete metric space and T a collection of real valued continuous functions on A. If
T is bounded at each point r A,
[)(r)[ `(r) < for all ) T.
then there is an open set l A and a constant ` < such that
[)(r)[ ` for all r l and ) T.
Proof. This result follows from the Baire Category Theorem of topology:
Theorem 19.4 (Baire Category Theorem). A complete metric space is not the union of
a countable number of nowhere dense sets.
Remark. Recall that the interior of a set o is the largest open set contained in o, that
is
int o = o
o
=
_
l o : l open.
and that a set o is nowhere dense if its closure o has empty interior. Thus a closed set is
nowhere dense if o o
c
. For the proof see Reed and Simon Chapter III or any book on
point set topology.
To prove the PUB, note that by assumption
A =
_
n
r : [)(r)[ : for all ) T.
Thus, at least one of the (closed) sets r : [)(r)[ : ) T has non-empty interior,
which is to say it contains an open set l. This is the open set claimed in the theorem.
Suppose A is a Banach space and each function ) T is sub-additive ()(r + )
)(r) + )()) and absolutely homogeneous ()(cr) = [c[)(r)). For instance each ) could be
of the form )(r) = [/(r)[ for some linear functional. Then
Theorem 19.5 (Principle of Uniform Boundedness for sub-additive functionals). Let A
be a Banach space and let T be a collection of real-valued continuous, sub-additive, absolutely
homogeneous functions on A. Suppose for each r A, [)(r)[ `(r) < for all ) T.
Then the function ) T are uniformly bounded in the sense that there is c < such that
[)(r)[[ c |r| for all r A and ) T.
Proof. Clearly the hypotheses of the PUB for metric spaces applies. Let l be the open
set claimed and let r
0
l. Since l is open there is c 0 such that || < c = r
0
+ l.
Now consider with || < c. We have, for ) T,
)() = )( +r
0
r
0
) )( +r
0
) +)(r
0
) 2`.
Thus for arbitrary r A and ) T,
)(r) =
2 |r|
c
)(
c
2 |r|
r)
4`
c
|r| .
19. WEAK CONVERGENCE 19-3
Corollary 19.6. Let A be a Banach space and let L be a collection of bounded linear
functionals that is pointwise bounded, so /(r) `(r) for all / L, then there is a constant
c < such that
|/| c for all / L.
Corollary 19.7. Let A be a Banach space and let o A be a weakly pre-compact
subset of A. Then there is a constant c < such that
|r| c for all r o.
In particular, any weakly convergent sequence is norm bounded.
Remark. Recall that a set o is pre-compact if o is compact. Thus any sequence in a
pre-compact set has a convergent subsequence, but the limit may lie outside o.
Proof. Think of points of A as functions on A
t
. Since o is weakly compact /(r) must be
bounded for each / as r ranges over o (otherwise we could nd a weakly divergent sequence).
By the PUB there is a constant c such that |r| c for all r o.
A function ) : A R, A a topological space, is called lower semi-continuous if r :
)(r) t is open for each t R. Such a function satises )(r) liminf
n
)(r
n
) for any
convergent sequence r
n
r. (Note that for each c 0 the set : )() )(r) c is
open and thus eventually contains r
n
so liminf r
n
)(r) c.)
Theorem 19.8 (Weak lower semicontinuity of the norm). Let A be a Banach space. The
norm || is weakly lower semicontinuous. In particular, if r
n
r in A then
|r| liminf |r
n
| .
Remark. 1) This should remind you of Fatous lemma from measure theory. 2) We have
already seen that the norm is not continuous, since it may jump down in a limit.
Proof. Fix t 0. Let A
t
1
denote the unit ball / : |/| 1 in A
t
. Note that |r| t
if and only if there is a linear functional / A
t
1
with [/(r)[ t . Thus
|r| t =
X

1
r : /(r) t
is weakly open.
LECTURE 20
Weak sequential compactness, weak

convergence and the weak

topology
Definition 20.1. A subset C of a Banach space A is called weakly sequentially compact
if any sequence of pints in C has a weakly convergent subsequence, whose weak limit is in
C.
Recall that sequential compactness is, in general, a strictly weaker notion than compact-
ness. They are equivalent, however, in metric spaces. In the present context, A is metrizable
in the (A. A
t
) topology if and only if A
t
is separable.
Proposition 20.1. A weakly sequentially compact set is bounded
Proof. Use the PUB. Details left as an exercise.
Theorem 20.2. In a reexive Banach space A the closed unit ball is weakly sequentially
compact.
Remark. We will see that the unit ball is, in fact, weakly compact in a reexive Banach
space, and more generally in the dual of a Banach space. However, the proof of that result
is far less constructive than the following.
Proof. Let
n
be a sequence of points in the unit ball. Let 1 be the closed linear
span span
n
. Since A is reexive, it follows that 1 is reexive, since
Theorem 20.3 (Thm. 15 of Chapter 8 in Lax). Any closed subspace of a reexive space
is reexive.
Proof. See Lax.
Since 1 = 1
tt
is separable, it follows that 1
t
is separable. So 1
t
contains a dense
countable subset :
j
. Consider the array of scalars
c
i,j
= :
i
(
j
).
The i
th
row is bounded by |:
i
|. Starting with the rst row we pick a subsequence
j
(1)
k
so
that :
1
(
j
(1)
k
) converges. Rene this subsequence again and again to produce subsequences

j
(n)
k
for each : so that :
1
(
j
(n)
k
). . . . . :
n
(
j
(n)
k
) all converge. Now let
.
n
=
j
(n)
n
.
Clearly :
j
(.
n
) converges as : for each ,. By density of :
j
in 1
t
it follows that
lim
n
:(.
n
) = (:)
exists for each : 1
t
. This limit is clearly a linear functional of : and since
|:(.
n
)| |:| |.
n
| |:| .
20-1
20-2 20. WEAK SEQUENTIAL COMPACTNESS, WEAK

CONVERGENCE AND THE WEAK

TOPOLOGY
the linear functional is bounded. Since 1 is reexive, we see that there is 1 such that
(:) = :(). Since the restriction of / A
t
to 1 gives an element of 1
t
, we have .
n

in A.
We may also consider a weak topology on the dual A
t
of a Banach space. That is the
weak topology (A
t
. A). A sequence n
n
of linear functionals is said to be weak convergent
to n if
limn
n
(r) = n(r) for all r A.
also denoted
wk

lim
n
n
n
= n.
Weak

convergence of measures is also known as vague convergence. If A is reexive then


weak

convergence is the same as weak convergence, but in general the weak

topology is
strictly weaker than the weak topology since the latter makes all linear functionals in A
tt
continuous.
Theorem 20.4. A weak

convergent sequence n
n
is uniformly bounded and
|n| liminf |n
n
| .
if n = wk

limn
n
.
Proof. Exercise.
Definition 20.2. A subset C of a dual Banach space A
t
is weak

sequentially compact
if every sequence of points in C has a weak

convergent subsequence with weak

limit in C.
Theorem 20.5 (Helly 1912). Let A be a separable Banach space. Then the closed unit
ball in A
t
is weak

sequentially compact.
Proof. Given n
n
A
t
with |n
n
| 1 and a countable dense subset r
n
of A, we can
use the diagonal process to select a subsequence
n
of n
n
so that
lim
n

n
(r
k
)
exists for every r
k
. By density of r
k
this extends to all of A:
lim
n

n
(r) = (r)
for all r A. One readily veries that is linear and bounded, so it is the desired limit.
In fact, more is true. The unit ball in A
t
is weak

compact, even if A is no separable:


Theorem 20.6 (Alaoglu). Let A
t
be the dual of a Banach space A. The unit ball of A
t
is wk

compact.
Proof. Let 1 be the unit ball in A
t
. Let 1 be the (uncountable) product space:
1 =

xX
1
x
. 1
x
= [|r| . |r|].
By the Tychonov theorem 1 is compact in the product topology. To complete the proof, we
embed 1 as a closed subset of 1.
The innite product space 1 is the collection of all functions 1 : A R such that
1(r) 1
x
for all r. Given / 1, [/(r)[ |/| |r| |r| so /(r) 1
x
for every r. Thus
1 1.
20. WEAK SEQUENTIAL COMPACTNESS, WEAK

CONVERGENCE AND THE WEAK

TOPOLOGY 20-3
Now the product topology on 1 is just the weakest topology such that coordinate eval-
uation 1 1(r) is continuous for every r. The restriction of this topology to 1 is just the
wk

topology on 1.
Thus we have embedded 1 as a subset of the compact space 1. It suces to show that
1 is closed. For each r. A and t R, let

x,y;t
(1) = 1(r +t) 1(r) 1().
a continuous map of 1 into the eld of scalars. Clearly 1
1
x,y;t
(0) and
1
x,y;
(0) is a
closed set. Thus
1

x,y,t

1
x,y;
(0).
so every element of 1 is linear. Since any 1 1 is also bounded by |r|, [1(r)[ |r|, we
conclude that 1 = 1.
Clearly Alaoglus theorem implies Hellys theorem. However, the proof of Hellys theorem
is much more useful. Often times what one really wants is to nd a convergent sequence.
The proof Hellys theorem gives you an idea how to construct it; Alaoglus theorem just tells
you it is there.
Corollary 20.7. The unit ball in a reexive space is weakly compact.
Remark. In fact weak compactness of the unit ball is equivalent to reexivity, a result
due to Eberlein (1947) and Smulyan (1940).
In particular, this result applies to any Hilbert space and to 1
p
, 1 < j < . The unit
ball in 1

is weak

compact since 1

= (1
1
)
t
. The unit ball in 1
1
is not weakly compact.
Here is what happens in 1
1
. Consider, for example, 1
1
([0. 1]), and let
)
n
(r) = :
[0,
1
n
]
(r).
So |)
n
|
L
1
= 1 and for any continuous function p C([0. 1]) T
_
1
0
p(r))
n
(r)dr p(0).T Thus
wk

lim)
n
dr = (r)dr in `([0. 1]) but )
n
has no weak limit in 1
1
. Of course, the sequence
)
n
has a weak

convergent subsequence in 1

([0. 1])
t
, which shows the existence of a linear
functional on 1

that restricts to p p(0) for continuous functions p. (We could have used
the Hahn Banach theorem to get this.)
Here is another example. On 1
1
([0. )) let
)
n
(r) =
1
:

[0,n]
(r).
Again |)
n
|
L
1
= 1. As measures )
n
dr 0 in `
0
([0. )), that is
_

0
)
n
(r)p(r)dr 0 p C
0
([0. )).
however, )
n
does not converge weakly to zero in 1

. Indeed for the constant function p 1,


_

0
)
n
(r)p(r) = 1.
LECTURE 21
An application: positive harmonic functions
Reading: 11.6 in Lax
Positive harmonic functions
We may apply weak

compactness to prove the following:


Theorem 21.1 (Herglotz). Let n be a function on the open unit disk 1 = [.[ < 1 such
that
(1) n(.) 0 for all . 1
(2) n is harmonic in 1, that is it satises the mean value property
n(.) =
1

2
_
D(z)
n(u)d:(u).
for all . 1 and < 1 [.[. Here : is Lebesgue measure on the disk and 1

(.)
is the disk of radius centered at ..
Then there is a unique nite, non-negative, Borel measure j on 1 = [.[ = 1 such that
n(.) =
_
D
1 [.[
2
[. u[
2
dj(u). (-)
Conversely, any such function is a non-negative Harmonic function on the disk.
Remark. The theorem implies [n(.)[ const.,(1 [.[), so non-negative harmonic
functions cannot blow up arbitrarily at 1. One might wonder if a similar theorem holds
for, say, real valued Harmonic functions. That is, given n Harmonic and real valued does
there exist a signed measure j such that n is the Poisson integral of j? A moments thought
shows that the answer is No!, for it is easy to construct a real valued harmonic function
which violates the estimate [n(.)[ const.,(1 [.[). For example, n(.) = Re 1,(1 .)
2
.
Proof. Note that n is continuous. To see this, rst observe that the mean value property
implies that n is locally integrable. Next, observe that
n(. +/) n(.) =
1

2
__
D(z+h)
n(u)d:(u)
_
D(z)
n(u)d:(u)
_
.
By dominated convergence the integral on the r.h.s. converges to zero as / converges to zero.
Since n is continuous, we may dierentiate
:
2
n(0) =
_
Dr(0)
n(.)d:(.) =
_
r
0
_
2
0
n(:e
i
)d:d:
with respect : and conclude that for every :
n(0) =
1
2
_
2
0
n(:e
i
)d.
21-1
21-2 21. AN APPLICATION: POSITIVE HARMONIC FUNCTIONS
Thus for each : (0. 1) the measure dj
r
() = (2)
1
n(:e
i
)d on the circle 1 has mass
n(0). Think of these measures as elements of the dual to C(1). By Hellys theorem, we
may nd a weak

convergent subsequence j
rn
. That is, there is a Borel measure j C(1)
t
such that
_
D
)()dj() = lim
n
1
2
_
D
)()n(:e
i
)d
for every ) C(1).
To complete the proof, we will use the identity
n(.) =
1
2
_
2
0
:
2
[.[
2
[:e
i
.[
2
n(:e
i
)d. (--)
valid for . 1
r
(0). Let us defer the proof for the moment and show how (--) implies the
representation (-). Fix . and let
)
r,z
(e
i
) =
:
2
[.[
2
[:e
i
.[
2
.
It is easy to see that )
r,z
)
z
uniformly as : 1, where
)
z
(e
i
) =
1 [.[
2
[e
i
.[
2
.
Thus the weak

convergence j
rn

_
j then implies
n(.) = lim
n
_
D
)
rn;z
(e
i
)dj
rn
() =
_
2
0
)
z
(e
i
)dj().
The identity (-) is a classical formula, which may be veried in a number of ways. One
of these is as follows. Let (.) denote the integral on the r.h.s. It is easy to show that is
harmonic in 1
r
(0) for this it suces to show that ([u[
2
[.[
2
),[u .[
2
is harmonic in
1
[w[
(0) for xed u. Furthermore, it is not too hard to show that
lim
sr
(:e
i
) = n(:e
i
).
since for any continuous function ) on the circle
1
2
lim
sr
_
2
0
:
2
:
2
[:e
i
:e
i
[
2
)(e
i
)d = )(e
i
).
(Exercise: verify this formula.) Thus n(.)(.) is a harmonic function on 1
rn
(0), continuous
up to the boundary and identically equal to zero there. It follows from the maximum
principle, applied to n and n, that n = 0 throughout. (The maximum principle
is a straightforward consequence of the mean value property and continuity.)
Herglotz-Riesz Theorem
An important application of the above is the following:
Theorem 21.2 (Herglotz-Riesz). Let 1 be an analytic function in the unit disk 1 such
that Re 1 0 in 1. Then there is a unique non-negative, nite, Borel measure j on 1
such that
1(.) =
_
2
0
e
i
+.
e
i
.
j(d) + i Im1(0).
HERGLOTZ-RIESZ THEOREM 21-3
Conversely every analytic function in the disk with positive real part can be written in this
form.
Proof. First apply the Herglotz theorem to Re 1. Let
G(.) =
_
2
0
e
i
+.
e
i
.
j(d).
So G and 1 are analytic functions on the disk whose real parts agree. It follows that 1 G
is constant and imaginary. However G(0) = Re 1(0) so 1(.) G(.) = i Im1(0).
The theorem is often used in the following form
Theorem 21.3. Let 1 be an analytic map from the upper half plane . : Im. 0
into itself. Then there is a unique non-negative Borel measure j on R and a non-negative
number 0 such that
_
R
1
1 +r
2
dj(r) <
and
1(.) = . + Re 1(i) +
_
R
1 +r.
r .
1
1 +r
2
dj(r). (- - -)
Furthermore
= lim
z
1(.)
.
.
and
dj(r) = wk

lim
y0
1

Im1(r + i)dr.
If lim
z
(1(.) .) = 1 exists and is real, and if lim
z
.(1(.) . 1) exists then j
is a nite measure and
1(.) = . +1 +
_
R
1
r .
dj(r).
Remark. Note that
1 +r.
r .
1
1 +r
2
=
1
r .
Re
1
r i
.
Proof. Consider the function
G() = i1
_
i
1
+ 1
_
.
This is an analytic map from the disk into the right half plane. By the Herglotz-Riesz
theorem
1
_
i
1
+ 1
_
= i
_
2
0
e
i
+
e
i

d() + Re 1(i)
Now let . = i(1 ),( + 1), so = (1 + i.),(1 i.) and
1(.) = i
_
2
0
e
i
(1 i.) + 1 + i.
e
i
(1 i.) 1 i.
d() + Re 1(i).
Now we dene a map : 1 1 R via
(e
i
) = i
1 e
i
1 + e
i
.
21-4 21. AN APPLICATION: POSITIVE HARMONIC FUNCTIONS
and let j = ;, that is
_
)d j =
_
) d
for functions ) C
0
(R). Now given p C(1), p p(1)1 vanishes at 1 and may be
written as
p p(1)1 = )
with ) = (p p(1)1)
1
. Thus
_
D
pd =
_
(p p(1)1)
1
d j +p(1)(1).
Since (1) = Im1(i), we conclude that
1(.) = Re 1(i) +. Im1(i) +
_
R
_
i
(1 + ir)(1 i.) + (1 ir)(1 + i.)
(1 + ir)(1 i.) (1 ir)(1 + i.)
.
_
d j(r).
since
1
(r) = (1 + ir),(1 ir). After simplifying, this gives
1(.) = . + Re 1(i) +
_
R
1 +r.
r .
d j(r).
with = Im1(i) j(R). The representation (- - -) follows with dj(r) = (1 +r
2
)d j(r).
The identity
= lim
z
1(.)
.
holds since
1
.
_
R
1 +r.
r .
d j(r) 0.
Furthermore, if lim
z
.(1(.) . 1) exists for some real number 1 then in particular
lim
t
t(Im1(it) it) = lim
t
t
_
R
Im
1 + itr
r it
d j(r)
exists and is nite. The integrand on the r.h.s. is
t
2
r
2
+t
2
+t
2
r
2
t
2
+r
2
=
t
2
r
2
+t
2
(1 +r
2
)
converges pointwise, monotonically to (1 +r
2
). Thus j is a nite measure, and
1(.) = . + Re 1(i) +
_
R
1
r .
dj(r) Re
_
R
1
r i
dj(r).
One checks now that
lim
z
(1(.) .) = Re 1(i) Re
_
R
1
r i
dj(r).
Presentation topics
Possible presentation topics
Please choose a presentation topic and discuss it with me by March 31. Presentations
will be 20 minutes in length and given in class April 16, 18, 21, 23 and 25.
(1) 9.1 Completeness of weighted powers in C
0
(R).
(2) 9.2 M untz Approximation Theorem.
(3) 11.2 Divergence of Fourier Series.
(4) 11.3 Approximate quadrature.
(5) 11.5 Existence of solutions to P.D.E.s
(6) 14.3 Completely monotone functions
(7) 14.7 Theorems of Caratheodory and Bochner
(8) 16.3.2 Hilbert Transform
(9) 16.3.3 Laplace Transform
(10) 16.4 Solution operators for hyperbolic equations
(11) 16.5 Solution operator for the heat equation
(12) 22.4 Operators dened by parabolic equations
Homework II
Homework II
Due: March 31, 2008
(1) (Ex. 1, Ch.6) Show that a norm that satises the parallelogram law comes from an
inner product.
(2) (Ex. 3, Ch. 6) Show that /
2
is complete.
(3) Let A be a reexive Banach space and 1 A a closed subspace of A. Show that
(1

= 1 .
(4) (Ex. 5, Ch. 8) Let A. j be a measure space with j a positive measure. Show that
if j(A) = 1 then |)|
p
is an increasing function of j
(5) Prove Theorem 11.3 of these notes.
(6) Prove that [r[
2d
= (d 2)[o
d1
[(r) in the sense of distributions on R
d
, d 3,
where = Laplacian, [o
d1
[ is the area of the unit sphere [r[ = 1 : r R
d
and
(r) is the Dirac delta function.
(7) Consider the map 1 dened on Schwarz functions on the real line by the principle
value integral:
1() = P.V.
_
R
1
r
(r)dr = lim
0
__

1
r
(r)dr +
_

1
r
(r)dr.
_
Show that 1 is a tempered distribution. Convolution with 1 is the Hilbert trans-
form.
(8) (Thm. 20.4 of the notes) Let A
t
be the dual of a Banach space. Show that / |/|
is weak

lower semi-continuous and that


|/| liminf
n
|/
n
| .
(9) Prove Prop. 20.1 of these notes.
(10) Show that the unit ball in an innite dimensional Hilbert space contains innitely
many disjoint translates of a ball of radius

2,4. Conclude that there is no non-


trivial translation invariant measure on an innite dimensional Hilbert space.
(11) A subset o of a Banach space is weakly bounded if for all / A
t
, sup
xS
[/(r)[ <
. Prove that o is weakly bounded if and only if it is strongly bounded, i.e.,
sup
xS
|r| < .
21-2
(12) Let A be a locally convex space.
(a) Let l be a compact neighborhood of 0. Show that one can nd r
1
. . . . . r
n
so
that l
n
i=1
(r
i
+
1
2
l), and thus a nite dimensional linear space 1 A with
l 1 +
1
2
l.
(b) Prove that l 1 + (
1
2
)
m
l for any :.
(c) Prove that l 1 = 1 .
(d) Conclude that A = 1 = 1 is nite dimensional.
Thus, any locally compact, locally convex, space is nite dimensional.
(13) Let A be a reexive Banach space. Is the open unit ball in A open in the weak
topology?
Part 6
Convexity
LECTURE 22
Convex sets in a Banach space
Reading: 8.4 and Ch. 12 of Lax
Definition 22.1. The support function o
M
: A
t
R of a subset ` of a Banach space
A over R is the function
o
M
(/) = sup
yM
/().
The support function o
M
of a set ` has the following properties:
(1) Subadditivity, o
M
(/ +:) o
M
(/) +o
M
(:).
(2) o
M
(0) = 0.
(3) Positive homogeneity, o
M
(c/) = co
M
(/) for c 0.
(4) Monotonicity: for ` `, o
M
(/) o
N
(/).
(5) Additivity: o
M+N
= o
M
+o
N
. (Recall that ` +` = r +[r ` and `.)
(6) o
M
(/) = o
M
(/)
(7) o
M
= o
M
.
The proof of these is left as an exercise.
In addition, we have
(9) o

M
= o
M
,
where
Definition 22.2. The closed convex hull of a subset ` of a Banach space, denoted

`
is the smallest closed convex set containing `.
Remark.

` is also the closure of the convex hull

`, where the convex hull is the
smallest convex set containing `. (Exercise)
Let us prove property (9), assuming the other properties. First by (8) it suces to show
o

M
= o
M
. Since `

`, by (5) we have o
M
o

M
. However,

` =
n

j=1
c
j
r
j
: r
j
`. c
j
0. and

j
c
j
= 1.
(Exercise: prove that this is in fact the smallest convex set containing `.) So for any point
in

` we have
/(
n

j=1
c
j
r
j
) =
n

j=1
c
j
/(r
j
) o
M
(/).
Thus o

M
o
M
.
Here are some examples:
i. If ` = r
0
, o
M
is just evaluation at r
0
.
ii. If ` = 1
R
(0) then o
M
(/) = 1|/| .
22-1
22-2 22. CONVEX SETS IN A BANACH SPACE
iii. If ` = 1
R
(r
0
) then ` = r
0
+1
R
(0) so o
M
(/) = 1|/| +/(r
0
).
iv. If ` is a closed subspace then o
M
(/) = 0 for / `

and for all other /.


Note that in the last example the set ` is unbounded. For bounded sets, o
M
: A
t
R,
however in general we dene o
M
as a map from A
t
R. We extend the order relation
and arithmetic to R by r and r + = for all r and c = for c 0.
This extended function satises all of the above properties.
If ` is bounded, then o
M
(/) const. |/| and is therefore continuous in the norm
topology, since by sub-additivity
[o
M
(/) o
M
(/
t
)[ maxo
M
(/ /
t
). o
M
(/
t
/) const. |/ /
t
| .
This fails if ` is unbounded, and also in the weak

topology. Nonetheless, we have the


following additional property
(10) o
M
is weak

lower semi-continuous
Indeed, since it is a sup of weak

continuous functions, we have


/ : o
K
(/) t =
_
zK
/ : /(.) t .
(Weak

continuity of a R valued function is dened in the same way as for a R valued


function.)
Theorem 22.1. The closed convex hull

` of a subset ` of a Banach space A over R
is equal to

` = . : /(.) o
M
(/) for all /.
Proof. Since o
M
= o

M
it follows that /(.) o
M
(/) for all .

`.
Now, suppose . ,

`. Since

` is closed there is an open ball 1
R
(.) with 1
R
(.)

` = .
By the geometric Hahn-Banach theorem we can nd a linear functional /
0
and c R such
that
/
0
(n) c < /
0
() for all n

` and 1
R
(.).
In particular, if |r| 1 then
/
0
(r) = /
0
(r +.) +/
0
(.) /(.) c .
so |/
0
| 1
1
(/
0
(.) c). Thus /
0
is bounded. From the denition of the norm of a linear
functional, we have
inf
|x|<R
/
0
(r) = 1|/
0
| .
Since for . +r 1
R
(.) we have c /
0
(. +r), we nd that
/
0
(.) c +1|/
0
| .
It follows that
/
0
(.) o
M
(/
0
) +1|/
0
| .
Thus /
0
is a linear functional such that /
0
(r) o
M
(/
0
).
The theorem shows that a closed, convex set 1 is specied as the set
1 = . :
K
(.) 0
where

K
(.) = sup
:||1
[/(.) o
K
(/)].
22. CONVEX SETS IN A BANACH SPACE 22-3
Since o
K
: A
t
R , the function
K
is initially dened as a map A R .
However, note that (.) = for some . if and only if o
K
(/) = for all /, in which case
(.) = for all . and 1 = A. For any proper closed, convex set 1 there is some / such
that o
K
(/) < and
K
: A R.
Since o
K
is weak

lower semi-continuous, it follows that, for xed ., /(.)o


K
(/) is weak

upper semi-continuous, that is for each t


/ : /(.) o
K
(/) < t
is weak

open. This observation is relevant, since |/| 1 is compact, and


Proposition 22.2. Let 1 be a compact topological space and let 1 : 1 R be
upper semi-continuous. Then 1 is bounded from above and attains its maximum.
Remark. We did this for lower semi-continuous functions, but it didnt make it into the
notes, so lets prove it again.
Proof. The sets 1(r) < t are increasing, open, and cover 1. By compactness 1
1(r) < t for some t. So 1 is bounded from above. Now let t
m
= sup
xK
1(r). Suppose
1(r) < t
m
for all r. Then the sets 1(r) < t for t < t
m
cover A. By compactness there is
then some t < t
m
such that 1 1(r) < t, contradicting t
m
= sup
xK
1(r). So there is a
point r
m
such that t
m
= 1(r
m
).
It follows that,

K
(.) = max
||1
[/(.) o
K
(/)].
The function
K
(.), being a sup of weakly continuous functions, is in turn weakly lower
semi-continuous. In particular,
1 = . :
K
(.) 0
is weakly closed! Thus we have obtained the following theorem:
Theorem 22.3 (Theorem 2, 12 of Lax). A convex set 1 of a Banach space is closed in
the norm topology if and only if it is closed in the weak topology.
This theorem is astounding, since there are certainly strongly closed sets that are not
weakly closed. (Weakly closed = strongly closed for any set.) For instance the complement
of an open ball r : |r| 1 is norm closed but not weakly closed. (Exercise: prove this.)
LECTURE 23
Convex sets in a Banach space (II)
Reading: 8.4
Recall that we showed last time that a closed convex set 1 in a Banach space is equal to
1 = r :
K
(r) 0 .
where

K
(r) = max
||1
[/(r) o
K
(/)]. o
K
(/) = sup
zK
/(.).
The weak lower semi-continuity of
K
showed that 1 is weakly closed. As a corollary we
have
Corollary 23.1. If A is reexive, then a bounded, norm closed, convex subset 1 is
weakly compact.
Remark. This may fail if A is not reexive. For instance in 1
1
(R) the set 1 of non-
negative functions 0 with integral
_
= 1 is convex, norm closed and bounded. However,
it is not weakly compact.
Similarly, this may fail if A is a dual and we take the weak

topology on A. For instance,


inside `
0
(R) the space of nite measures on R the space of probability measures is norm
closed, bounded, and convex but is not weak

closed.
All of the suggests that
K
(r) might be a decent measure of how far a point r is from
1. In fact, it is precisely the distance of r to 1!
Theorem 23.2. Let 1 be a closed, convex subset of a Banach space A. Then

K
(r) = inf
uK
|r n| .
Proof. Suppose r 1. Then /(r) o
K
(/) 0 for all / so the maximum is attained at
/ = 0 and
K
(r) = 0.
If r , 1 and n 1 and |/| 1, then
o
K
(/) /(n) = /(n r) +/(r) /(r) |n r| .
Thus
|n r| sup
||1
[/(r) o
K
(/)].
On the other hand, if 1 < inf
uK
|n r|, then the convex set 1+1
R
(0) still has positive
distance from r. Thus by the Theorem of last lecture there is /
0
A
t
such that
o
K
(/
0
) +1|/
0
| = o
K+B
R
(0)
(/
0
) < /
0
(r).
This inequality is homogeneous under positive scaling, so we may take |/
0
| = 1 to conclude
1 < /
0
(r) o
K
(/
0
) sup
||1
[/(r) o
K
(/)].
As 1 was any number less than inf
uK
|n r| the reverse inequality follows.
23-1
23-2 23. CONVEX SETS IN A BANACH SPACE (II)
Let us turn all of this around. Suppose we are given a function o : A
t
R .
Consider the set
1 = . A : /(.) o(/) for all / A
t
.
Then 1 is clearly convex and weakly closed, and its support function
o
K
(/) = sup
xK
/(.) o(/).
Can it happen that the inequality is strict? Of course it can, as the function o is arbitrary.
However, if we assume that o is, like o
K
, positive homogeneous, sub-additive, maps 0 to 0
and is weak

lower semi-continuous then the answer is no, at least if A is reexive.


Theorem 23.3. Let A be a reexive Banach space and let o : A
t
R be a weak

lower semi-continuous function which is positive homogeneous, sub-additive, and maps 0 to


0. Then
o(/) = sup
xK
/(r).
where
1 = r : /(r) o(/) for all /
.
Proof. To begin, let us prove the theorem under the additional restriction that o is
bounded: [o(/)[ |/|. Then 1 is clearly bounded, since r 1 = |r| . Since
o(0) = 0 the identity holds for / = 0. So x a non-zero linear functional /
0
. Let us dene a
linear functional 1 A
tt
, the double dual, via Hahn-Banach. Begin on the one-dimensional
subspace span/
0
and let
1(t/
0
) = to(/
0
).
By positive homogeneity and sub-additivity 1(t/
0
) o(t/
0
) for all t R. (Note that
o(/
0
) o(/
0
).) Thus by Hahn-Banach there is a linear functional 1 on A
t
which satises
1(/) o(/)
for every / A
t
. Since o(/) |/| this functional is bounded. As A is reexive there is
. A such that 1(/) = /(.). Since /(.) o(/) for all / this point . 1, and since
o(/
0
) = 1(/
0
) = /
0
(.).
we have
o(/
0
) = max
xK
/
0
(r).
As /
0
was arbitrary, this completes the proof for o bounded.
To extend this to unbounded o, note that
< inf
||1
o(/) 0.
by wk

lower semi-continuity. Let = inf


||1
o(/). Then by positive homogeneity,
o(/) |/| .
that is o is bounded from below.
Now, for each c dene
1

= r : /(r) o

(/) for all / .


23. CONVEX SETS IN A BANACH SPACE (II) 23-3
where
o

(/) = inf

1
,
2
X

:
1
+
2
=
_
o(/
1
) +
1
c
|/
2
|
_
.
It is left as an exercise to show, for each c 0, that o

is positive homogeneous, sub-additive


and maps 0 to 0. Note also that
|/| o

(/)
1
c
|/| . and o

(/) o(/).
so o

is bounded and smaller than o.


Now, in fact,
1

= 1 11

(0).
Indeed, given r 1 1
1/
(0) we have
/(r) = /
1
(r) +/
2
(r) o(/
1
) +
1
c
|/
2
|
if / = /
1
+ /
2
, so /(r) o

(/) and r 1

. On the other hand if r 1

then /(r) o(/)


and
1

|/| for all / so r 1 1


1/
(0).
It follows that
1 =

..
so
sup
xK
/(r) = sup

sup
xK
/(r) = sup

(/).
Thus, it suces to show, for xed /, that
o(/) = sup

(/).
To show this, note that o

increases as c decreases, so
o
0
(/) := lim
0
o

(/) = sup
>0
o

(/)
exists and (since o

o) satises
0 o
0
(/) o(/).
Furthermore, for each c we can nd /

such that
o

(/) o(/ /

) +
1
c
|/

| o

(/) +c.
Since o(/ /

) |/| |/

| we see that
|/| +
_
1
c

_
|/

| o

(/) +c.
Consider the following cases: (1) |/

| ,c is bounded as c 0 or (2) |/

| ,c is unbounded as
c 0 . In case (2) lim
0
o

(/) = o
0
(/) = o(/) = . On the other hand, in case (1) /

0
so by weak

lower semi-continuity of o we nd that


o(/) liminf
0
o(/ /

) o
0
(/) liminf
0
1
c
|/

| o
0
(/).
which completes the proof.
LECTURE 24
Krein-Milman and Stone-Weierstrass
Reading: 13.3
Definition 24.1. An extreme subset o of a convex set 1 is a subset o 1 such that
(1) o is non-empty and convex
(2) If r o and r = t + (1 t). with . . 1 then . . o.
An extreme point is a point r 1 such that r is an extreme subset.
The following is a classical result due to Caratheodory:
Theorem 24.1. Every compact convex subset 1 of R
N
has extreme points, and every
point of 1 can be written as a convex combination of (at most) ` + 1 extreme points.
Remark. The proof is left as an exercise. Use induction on `. The case ` = 1 is easy!
Theorem 24.2 (Krein and Milman). Let A be a locally convex space, 1 a non-empty,
compact, convex subset of A. Then
(1) 1 has at least one extreme point
(2) 1 is the closure of the convex hull of its extreme points.
Proof. Consider the collection c of all nonempty closed extreme subsets of 1. Since
1 c this collection is nonempty. Partially order c by inclusion. We wish to apply Zorns
lemma to see that c has a maximal element, i.e., a set that is minimal with respect to
conclusion.
Let T c be a totally ordered sub-collection. That is T = 1

: with some
totally ordered index set and 1

if . Clearly T is a candidate for an upper


bound. To see this we must show that T is nonempty, closed, and extreme.
Clearly T is closed. Furthermore,one easily shows that the intersection of an arbitrary
family of extreme sets is extreme provided it is non-empty. Thus we need only show T is
non-empty. Here we use compactness of 1 in a crucial way.
Recall that a family T of closed sets is said to have the nite intersection property (FIP)
if any nite collection 1
1
. . . . 1
n
T has non-empty intersection: 1
1
1
n
,= , and
there is the following result connecting the FIP and compactness:
Theorem 24.3. A topological space ` is compact if and only if every collection T of
closed sets with the FIP satises
T ,= .
Proof. Suppose ` is compact and T = . Then | = ` with / = 1
c
: 1 T.
Thus ` =
n
j=1
1
c
j
for some nite collection. Thus
n
j=1
1
j
= and T does not have the FIP.
Conversely, if T is a collection of closed sets with the FIP and nonetheless T = , then
| = 1
c
: 1 is an open cover of ` with no nite subcover so ` is not compact.
24-1
24-2 24. KREIN-MILMAN AND STONE-WEIERSTRASS
Regarding our collection T , it clearly has the FIP since it is totally ordered, so any nite
collection 1
1
. . . . 1
n
T has a minimal element. Thus T , = .
Sp c has minimal elements. We claim that any such minimal element is a one point set.
Indeed, suppose 1 c contains two distinct points (it must also contain the line segment
joining them). Then there is a continuous linear functional / on A that separates these
points. Let ` 1 be the set
` = r 1 : /(r) = max
zE
/(r).
Then ` is a non-empty, proper, closed subset of 1. It is clearly convex and one easily shows
it is an extreme subset of 1. It follows that ` is an extreme subset of 1 (why?) and ` 1
so 1 is not minimal.
This proves (1): 1 has at least one extreme point. (It might have only one: 1 could be
the set x.) Since any closed extreme subset 1 of 1 is itself a closed convex set we nd
that every extreme subset of 1 has an extreme point r. It is easy to see that an extreme
point of 1 is also an extreme point of 1 (since 1 is an extreme subset). Thus
Every closed, extreme subset of 1 contains an extreme point of 1.
Let 1
e
denote the set of extreme points,

1
e
its convex hull, and

1
e
its closed convex
hull which is the closure of

1
e
. One easily shows that

1
e
is convex. (Exercise)
Clearly

1
e
1, so since 1 is closed

1
e
1. On the other hand if . ,

1
e
then
there is an open set l with . l 1
e
. We may take l to be convex. By the geometric
Hahn-Banach there is a linear functional / and c R such that
/(r) < c /() for all r l and

1
e
.
(As l is open, all points of l are interior, so the rst inequality is strict.) The gauge function
of l .,
j
Uz
(r) = inft : r,t l ..
is a continuous seminorm on A (why?). If r,t l . we have
1
t
/(r) = /(
r
t
+.) /(.) < c /(.).
Thus
/(r) (c /(.))j
Uz
(r).
and / is a continuous linear functional because
Lemma 24.4. A linear functional on locally convex space is continuous if and only if it
is bounded with respect to some continuous seminorm.
Proof. Exercise.
Since 1 is compact / achieves its minimum on 1. Let 1 be the set of minimizers. Then
1 is closed, convex and extreme. (Why?) By the above derived result 1 contains an extreme
point. Thus
min
xK
/(r) = min
xKe
/(r) /(.).
Thus . , 1.
An interesting application of this theorem is:
24. KREIN-MILMAN AND STONE-WEIERSTRASS 24-3
Theorem 24.5 (Stone-Wierstrass). Let o be a compact Hausdor space and C(o) the
set of real valued continuous functions on o. Let 1 C(o) be an sub-algebra, that is 1 is
linear subspace and ). p 1 = )p 1. If the constant function 1 1 and if for any
pair of points j. o there is a function ) 1 with )(j) ,= )(), then 1 is dense in C(o)
in the max norm.
This theorem is due to Stone and generalizes the classical result of Weierstrass on ap-
proximation of continuous functions on an interval with polynomials: o = [0. 1] and 1 =
polynomials.
Sketch of proof. Consider the collection ^ of all nite measures j on o such that
_
)dj = 0 for all ) 1. Then o is dense if and only if / = 0. (Why?)
By construction ^ is weak

closed. So 1 = ^ 1
1
(0) is weak

compact. Furthermore
it is convex. Suppose 1 contains a non-zero measure j. Then 1 must contain a non-zero
extreme point j. Since j is extreme we must have |j| = 1 (otherwise we could write j as
a linear combination of 0 and some multiple of j).
Suppose such a j exists. Since 1 is an algebra
_
)pdj = 0 for all ). p 1. Thus
pdj ^ for all p 1. Now let p be a continuous function on o with 0 < p(j) < 1 for all
j o. Let
c =
_
pd [j[ . / =
_
(1 p)d [j[ .
So c. / 0 and c +/ = 1 and pdj,c, (1 p)dj,/ 1. Since
dj = c
p
c
dj +/
(1 p)
/
dj
we must have pdj,c = dj (recall that j is an extreme point).
Consider the support of j:
supp j = j : [j[(l) 0 for any open neighborhood of j.
Since dj = pdj,c for any p 1 that satises 0 < p(j) < 1 we must have p c on supp j.
(Why?) Suppose j and are distinct points in o. Then there is a function p 1 such that
0 < p < 1 and p(j) ,= p(). (Just add a large constant to a function in 1 that separates j
and ). Thus at most one of the points j. lies in the support of j. That is the support of
j is a single point supp jj
0
! Since [j[(1) = |j| = 1 we have
_
)(j)dj(j) = )(j
0
) or
_
)(j)dj(j) = )(j
0
).
However, we have
_
1dj = 0 which is a contradiction, so ^ = 0 and 1 is dense.
Following the above proof, we nd
Theorem 24.6. The extreme points of the unit ball j :
_
d[j[ 1 /(o) are the
point masses (j j
0
).
Theorem 24.7. If C(o) is a proper closed sub-algebra that separates points of o
then = ) : )(j
0
) = 0 for some j
0
o.
The algebras
p
= ) : )(j) = 0 are exactly the maximal ideals of C(o). It is no
accident that the set of maximal ideals is the set of proper sub-algebras that separate points
and is in one to one correspondence with extreme points of the unit ball in C(o)
t
which is
in one to one correspondence with o.
LECTURE 25
Choquet type theorems
Reading: 13.4 and 14.10
In nite dimensions, Caratheodorys Theorem shows that points of a compact set may be
expressed as a convex combination of extreme points: no more than ` +1 points are needed
in dimension `. (A simplex shows that this number is optimal.) The following generalizes
this idea to LCSs:
Theorem 25.1. Let A be an LCS and 1 a non-empty compact, convex subset of A. For
any n 1 there is a Borel probability measure j
u
on 1
e
the closure of the set of extreme
points such that
/(n) =
_
Ke
/(r)dj
u
(r) (-)
for all / A
t
.
Remark. The integral in the theorem is understood as the identity
n =
_
Ke
rdj
u
(r).
in the weak sense. This expresses n as a generalized convex combination of points of 1
e
.
Lax presents without proof a sharper theorem , due to Choquet, in which the measure j
u
is dened on 1
e
provided 1 is metrizable. Any representation of type (-) is called a Choquet
decomposition. Such representations need not be unique, even in nite dimensions, for
example if 1 is a disk in R
2
.
Proof. The general idea of the proof is to show that given a function ) on 1
e
we can
evaluate it at a pont n 1 and produce a bounded linear functional: ) )(n). We then
represent this linear functional as a measure on 1
e
via Riesz-Kakutani.
There are two complications: (1) not every function ) on 1
e
can be extended in a
reasonable, or unique way to n 1 and (2) Riesz-Kakutani does not apply since 1
e
may
not be compact. The second complication is easily dealt with by replacing 1
e
with its
closure 1
e
which, being a closed subset of a compact space (1) is compact.
As for the rst: any linear functional / A
t
can be evaluated at n. Furthermore, being
continuous and linear, / achieves its max and min over 1 on the set of extreme points 1
e
.
(We saw this in the proof of Krein-Millman.) That is
min
xKe
/(r) /(n) max
xKe
/(r) for all n 1.
It follows that if /
1
(r) = /
2
(r) for r 1
e
then /
1
= /
2
on 1, so any linear functional
is determined on 1 by its restriction to 1
e
. Let 1 C(1
e
) be the subspace of maps
) : 1
e
R with
)(r) = /(r)
25-1
25-2 25. CHOQUET TYPE THEOREMS
for some / A
t
, or )(r) = c for some c R. Dene a linear functional 1 on 1 by
1()) =
_
/(n) )(r) = /(r)
c )(r) = c.
Then 1 is a positive linear functional on 1 and 1 contains the constant functions. By
Theorem 3.1 of these notes, we may extend 1 as a positive linear functional on the space of
all functions. Restricting this linear functional to C(1
e
), we obtain by Riesz-Kakutani that
1()) =
_
Ke
)(r)dj
u
(r) for all ) C(1
e
)
for some j
u
. In particular (-) holds. Since j
u
(1) = 1(1) = 1 j
u
is a probability measure.
The Riesz-Kakutani theorem, which was used in the above proof, is it-self an example of
this theorem. Indeed, let 1 C(o)
t
be the closed unit ball of the dual of C(o) with o a
compact Hausdor space. We saw last time that the extreme points 1
e
are point evaluations:
/
p,1
()) = 1 )(j)
for some j o. Thus 1
e
is in one to one correspondence with o 1. +1. Exercise:
show that 1
e
= o 1. +1 as topological spaces, where 1
e
has the weak

topology and
o 1. +1 has the product topology using the discrete topology on 1. +1 (all sets
all 4 of them are open). The above theorem shows that any linear functional / 1, with
norm no larger than 1, can be written as a combination
/()) =
_
S1
/
p,
())d:(j. ) =
_
S
)(j)d:(j. +) )(j)d:(j. 1) =
_
S
)(j)dj(j)
with dj(j) = d:(j. +) d:(j. 1).
Measure preserving maps
Let be a compact metric space and let 1 : be a homeomorphism. Consider the
set 1 of all probability measures on that are invariant with respect to 1:
1 = j /() : j 0. j() = 1. and j(o) = j(1(o)) for all measurable o .
We say that j is ergodic under 1 provided 1(o) o = j(o) = 0 or 1.
The following is one of the most important applications of Choquet type theorems.
Theorem 25.2. The set 1 is non-empty, convex and compact in the weak

topology.
The extreme points of 1 (which exist by Krein Millman) are the ergodic measures. Every
invariant measure can be represented uniquely as an average of ergodic measures.
Sketch of proof. Convexity of 1 is easy. Since 1 is clearly a subset of the unit ball
in /() it is compact once it is closed. Weak

closure follows since


j 1
_

)(1
1
())dj() =
_

)()dj()
for all ) C(). To see that 1 is non-empty, pick a point
0
and consider the sequence
of measures
n
given by
_
)()d
n
() =
1
:
n

j=0
)(1
j
(
0
)).
MEASURE PRESERVING MAPS 25-3
This is a sequence of probability measures on . Since the unit ball is weak

compact there
is a weak

convergent subsequence. The limit of this subsequence satises 0, () = 1,


and
_
)(1
1
())d() = lim
k
1
:
k
n
k

j=0
)(1
j1
(
0
))
= lim
k
1
:
k
n
k

j=0
)(1
j
(
0
)
1
:
k
)(
0
) +
1
:
k
)(1
n
k
1
(
0
)) =
_
)()d().
Suppose j 1 is not ergodic. Then there must be a disjoint union
1

2
= with
1(
j
)
j
and j(
j
) 0. Let j
j
be the restriction of j to
j
renormalized to be a
probability measure:
j
j
(o) =
j(o
j
)
j(
j
)
.
Then each j
j
is invariant under 1 and
j = j(
1
)j
1
+j(
2
)j
2
so j is not extreme.
Conversely, suppose j is not extreme, so
j = c:
1
+ (1 c):
2
with :
1
,= :
2
1 and 0 < c < 1. Suppose :
2
<< :
1
. Then by Radon-Nikodym there is
) 1
1
(:
1
), ) , 1, such that
dj = (c + (1 c)))d:
1
.
Since j and :
1
are invariant under 1 we must have )(1()) = )() for :
1
almost every .
Since ) , 1 the sets
1
= : )() 1 and
2
= : )() 1 are separately invariant
under 1 and both have positive j measure. So j is not ergodic.
If :
2
,<< :
1
then the situation is even happier. For then there is some set o such that
:
1
(o) = 0 but :
2
(o) 0. Let
1
=

j=0
1
j
(o). Then :
1
(
1
) = 0, :
2
(
1
) :
2
(o) 0
and
1
is invariant under 1. Since :
1
(
1
) = 0 we must have j(
1
) 1 c so j(
2
) 0
with
2
=
1
. Thus j is not ergodic.
The representation of j as a unique integral over ergodic measures is clearly a Choquet
type representation we need not take the closure of extreme points since the space of
measures is metrizable in the weak

topology as it is the dual of a separable Banach space


(this was Choquets theorem which we did not prove). The uniqueness requires a separate
argument, for which Lax refers a paper of Oxtoby in Bull. AMS 58 (1952).
Part 7
Bounded Linear Maps
LECTURE 26
Bounded Linear Maps
Definition 26.1. Let A and 1 be Banach spaces. A linear map ` : A 1 is bounded
if there is some 0 c < such that
|`r|
Y
c |r|
X
.
The smallest such c is called the norm, or operator norm, of `, denoted |`|.
Theorem 26.1. A linear map ` : A 1 is continuous if and only if it is bounded.
Proof. Bounded implies Lipschitz continuous since
dist(`r. `) = |`r `| |`| |r | = |`| dist(r. ).
On the other hand if ` is not bounded then there is a sequence r
n
such that
|`r
n
| :|r
n
| .
As this inequality is invariant under scaling we may take |r
n
| = 1,

: so r
n
0 but
|`r
n
| .
Theorem 26.2. The operator norm has the following properties
|c`| = [c[ |`| for all c 1 (homogeneity)
|`| 0 and |`| = 0 if and only ` 0 (positivity)
|` +1| |`| +|1| (sub-additivity)
That is the operator norm is a norm.
Note that the norm is dened to be
|`| = sup
x,=0
|`r|
|r|
.
By homogeneity this is the same as
|`| = sup
|x|=1
|`r| .
If ` is bounded on a normed space which is incomplete then ` has an extension to the
completion (as does any continuous function) which is also bounded, with the same norm.
Definition 26.2. Let L(A. 1 ) denote the set of all bounded linear maps from A to 1 .
Theorem 26.3. Let A be a normed space and let 1 be a Banach space. Then L(A. 1 )
is a Banach space under the operator norm.
Proof. Since || is a norm, we need to show completeness. Suppose `
n
is a Cauchy se-
quence. For each r A it follows that `
n
r is a Cauchy sequence in 1 , since |`
n
r `
m
r|
26-1
26-2 26. BOUNDED LINEAR MAPS
|`
n
`
m
| |r|. By completeness of 1 there is a limit. Call this limit `r. It is easy to see
that ` is linear. Moreover, if |r| = 1,
|`r| = lim
n
|`
n
r| limsup
n
|`
n
| < .
so |`| is nite and ` is bounded.
The operator norm topology on L(A. 1 ) is called the uniform topology. There are two
other natural topologies on this space
Definition 26.3. The strong topology on L(A. 1 ) is the weakest TVS topology such
that all functions ` `r are continuous from L(A. 1 ) 1 . The weak topology (or
weak operator topology) on L(A. 1 ) is the weakest TVS topology such that all maps `
/(`(r)), for / 1
t
and r A, are continuous.
Weak and strong sequential convergence are dened similarly. A sequence `
n
converges
strongly if
: lim
n
`
n
r exists in 1 for every r.
and converges weakly if
wk lim
n
`
n
r exists in 1 for every r.
Note that the weak operator topology on L(A. 1 ) is potentially much weaker than the weak
topology on L(A. 1 ) as a Banach space: the linear functionals ` /(`(r)) are just one
kind of linear functional on L(A. 1 ).
Definition 26.4. The transpose of a linear operator ` L(A. 1 ) is the map `
t

t
. A
t
dened by
r. `
t
/) = `r. /) .
where we use the dual pairing notation . /) = /().
Definition 26.5. The null space of a linear operator ` is the set
`
M
= r A : `r = 0.
The range of ` is the set
1
M
= `r : r A.
Theorem 26.4. Let ` L(A. 1 ) with A and 1 normed spaces. Then
(1) `
t
is bounded and |`
t
| = |`|
(2) The nullspace of `
t
is the annihilator of the range of `:
`
M
= 1

M
.
(3) The nullspace of ` is the annihilator of the range of `
t
:
`
M
= 1

M
= r : r. /) = 0 if / 1
M
.
(4) (c` +/`)
t
= c`
t
+/`
t
Proof. Exercise, or see Lax.
As a corollary we see that `
M
and `
M
are weak and weak

closed respectively. Since


these sets are subspaces this is the same as being closed. This is the rst part of
Theorem 26.5. Let ` L(A. 1 ) with A and 1 normed spaces. Then
26. BOUNDED LINEAR MAPS 26-3
(1) `
M
is a closed linear subspace of A.
(2) `, regarded as a map
`
0
:
A
`
M
1
is one-to-one, bounded, with |`
0
| = |`| and 1
M
0
= 1
M
.
Proof. The facts that `
0
is one-to-one and 1
M
0
= 1
M
are general facts about linear
maps. To see that `
0
is bounded, recall that A,`
M
is the set of equivalence classes [r] with
r if r `
M
and we put a norm on this space by
|[r]| = inf
yN
M
|r +| .
Since `
0
[r] = `r = `(r +) for any `
M
we have
|`| = sup
x,=0
|`r|
|r|
= sup
x,=0
sup
yN
M
|`(r +)|
|r +|
= sup
[x],=0
|`
0
[r]|
|[r]|
= |`
0
| .

Regarding convergence of adjoints we have the following


Proposition 26.6. If wk lim`
n
= ` then wk lim`
t
n
= `
t
.
Proof. Exercise.
However, this does not hold for strong limits or uniform limits. For example, on any /
p
space, 1 < j < , let o
j
be the ,
th
forward shift, the map o
j
L(/
p
. /
p
) given by
o(c
0
. c
1
. . . .) = (0. . . . . 0
. .
j zeroes
. c
0
. c
1
. . . .).
Then o
j
converges to zero weakly, but
|o
j
a|
p
= |a|
p
.
so o
j
does not converge to zero strongly. On the other hand the adjoint o
t
j
is the backwards
, shift,
o
t
j
(c
0
. c
1
. . . .) = (c
j
. c
j+1
. . . .).
This map converges to zero strongly since
_
_
o
t
j
a
_
_

0
for 1 < j < . (It does not converge to zero strongly for j = 1.)
LECTURE 27
Principle of Uniform Boundedness and Open Mapping Theorem
Reading: 15.315.5
As for linear functionals, we have the following useful criteria for strong/weak conver-
gence:
Proposition 27.1. Let `
n
L(A. 1 ) be a sequence of bounded maps between Banach
spaces A and 1 . Suppose that `
n
are uniformly bounded:
sup
n
|`
n
| < .
(1) If `
n
r converges in norm for all r in a dense subset of A then `
n
converges in
the strong operator topology.
(2) If `
n
r converges weakly for all in a dense subset of A then `
n
converges in the
weak operator topology.
Proof. Exercise.
As in the case of linear functionals, boundedness turns out to be necessary for convergence
as well:
Theorem 27.2 (Principle of Uniform Boundedness). Let A and 1 be Banach spaces and
let / L(A. 1 ) be a collection of bounded linear maps. If for each r A and / 1
t
there
is a constant c(r. /) such that
[`r. /)[ c(r. /) for all ` /.
then / is uniformly bounded, i.e., there is c < such that
|`| c for all ` /.
Sketch of proof. This follows very closely the proof of the PUB for linear functionals,
Corollary 19.6 of these notes. First apply that result to conclude that for each r A there
is c(r) with |`r| c(r) for all ` /. Then consider the collection of real values,
continuous, positive homogeneous functions )(r) = |`r| on the Banach space A. Apply
Theorem 19.5 to conclude that |`r| c |r| for all ` /.
The PUB follows from the Baire category theorem. This theorem also implies several
results which show that bounded linear maps have a number of useful properties in addition
to continuity.
Theorem 27.3. Let A and 1 be Banach spaces and ` : A 1 a bounded linear map
of A onto 1 . Then the image of the unit ball `1
1
(0) contains an open ball around the
origin in 1 .
Corollary 27.4 (Open mapping theorem). A bounded linear map ` from A onto 1 ,
with A and 1 Banach spaces, is open that is `(l) is open in 1 for any open subset l A.
27-1
27-2 27. PRINCIPLE OF UNIFORM BOUNDEDNESS AND OPEN MAPPING THEOREM
Proof. Let l be open in A. Then given `(l) we have = `r for some r l.
Since l is open r+c1
1
(0) l for some c 0. Thus `r+c`1
1
(0) `l so `l contains
an open ball centered at = `r.
Proof of Theorem. Since ` maps A onto 1 we have
1 =

n=1
`1
n
(0).
By the Baire category theorem at least one of the sets `1
n
(0) is dense in some open set.
That is there are 0 and 1 such that `1
n
(0) ( +1

(0)) is dense in 1

(0). Since
` is onto, there is r A such that `r = . So `(1
n
(0) r) is dense in 1

(0). By the
triangle inequality,
1
n
(0) r 1
n+|x|
(0).
By homogeneity we conclude that for every : 0, `1
r
(0) is dense in 1
r
(0) with
= ,(: +|r|).
Now let 1

(0) 1 . Then there is r


1
1
1
(0) such that
| `r
1
|
1
2

Let
1
= `r
1
. So
1
11
2

(0). Thus there is r


2
11
2
(0) such that |
1
`r
2
|
1
4
.
That is
| `(r
1
+r
2
)|
1
4
.
Following this procedure construct, by induction, a sequence r
1
. r
2
. . . . with
(1) r
j
1 1
2
j1
(0) A
(2)
_
_
_ `

n
j=1
r
j
_
_
_
1
2
j
.
Now the partial sums

n
j=1
r
j
converge as : to some point r A with
|r|

j=1
2
1j
= 2.
Clearly = `r. Thus 1

(0) `1
2
(0) and the result follows by homogeneity.
Theorem 27.5. Let A and 1 be Banach spaces and ` : A 1 a bounded, one-to-one
map of A onto 1 . Then the inverse map `
1
is bounded from 1 A.
Note in particular that
|`r|
1
|`
1
|
|r| .
Thus bounded, one-to-one and onto = ` is bounded from below! This is somehow
amazing because we were not given any quantitative information on the inverse just its
existence.
Proof. Since `1
1
(0) 1
d
(0) for some d 0 we have
`
1
1
1
(0)
for all 1
d
(0). By homogeneity |`
1
| 1,d.
Definition 27.1. A map ` : A 1 is closed if whenever r
n
r in A and `r
n

in 1 then `r = .
27. PRINCIPLE OF UNIFORM BOUNDEDNESS AND OPEN MAPPING THEOREM 27-3
Theorem 27.6 (Closed Graph Theorem). Let A and 1 be Banach spaces and ` : A
1 a closed linear map. Then ` is bounded.
Proof. Let the graph G of ` by the set of all pairs (r. `r) with r A. By linearity
of `, G is a linear space under coordinate-wise addition and scalar multiplication. Give G
a norm by dening
|(r. `r)| = |r| +|`r| .
Since ` is closed, the space G is complete in this norm. Now let 1 : G A be the
map 1(r. `r) = r. So 1 is bounded (|1| 1), one-to-one and onto. It follows that
1
1
r = (r. `r) is bounded. Thus there is a constant c < such that
|(r. `r)| c |r| .
So
|`r| (c 1) |r| .
Definition 27.2. Let A be a linear space and let ||
1
and ||
2
be two norms on A. The
norms ||
j
. , = 1. 2 are called compatible if whenever a sequence converges in both norms
the limits are equal.
For example, on 1
1
1
2
the 1
1
and 1
2
norms are compatible.
Corollary 27.7. Let A be a linear space and let ||
j
,, = 1. 2 be compatible norms. If
A is complete in both norms then the two norms are equivalent: there is c (0. ) such
that c
1
|r|
1
|r|
2
c |r|
1
.
Proof. Let A
j
, , = 1. 2, be the Banach space which is A under norm ||
j
, , = 1. 2.
The identity map 1 : A
1
A
2
is closed by compatibility of the norms. Thus 1 is bounded.
Likewise 1 : A
2
A
1
is bounded.
LECTURE 28
The Spectrum of a Linear Map
Reading: 15.5 and 20.1
First some observations about composition of linear maps. If ` : A 1 and ` : 1 2
are linear maps their composition (``)r = `(`r) is a map `` : A 2.
Theorem 28.1. If ` and ` are bounded then so is ``. Furthermore
(1) |``| |`| |`|
(2) (``)
t
= `
t
`
t
.
Proof. Exercise, or see Lax.
We now consider some general properties of linear maps from a Banach space A into
itself. Let L(A) = L(A. A). A map ` L(A) is invertible if ` maps A onto A and is
11. We saw last time that it follows from the open mapping theorem that `
1
is bounded
as well. Let (L(A) denote the set of all invertible maps in L(A).
Proposition 28.2. If 1. 1 (L(A) then so is 11 and (11)
1
= 1
1
1
1
.
Theorem 28.3. (L(A) is an open set in the uniform topology.
Proof. We must show that if 1 is invertible then so is every map of the form 1 +
with || < c for some c 0. Since 1 + = 1(1 + 1
1
) and |1
1
| |1
1
| || it
suces to prove this for 1 = 1 the identity map.
The basic idea is to make use of the geometric series to write (1 +)
1
as

n=0
(1)
n

n
.
Since
|
n
| ||
n
this series converges once || < 1. But then
(1 +)
N

n=0
(1)
n

n
= 1 + (1)
N

N+1
1
as ` so (1 +) is invertible.
Definition 28.1. The resolvent set (`) of a linear operator ` is the set of C such
that (1 `) (L(A). The spectrum (`) of ` is the complement of (`), that is the
set of such that 1 ` is not invertible.
Theorem 28.4. The resolvent set (`) is open and contains the set : [[ |`|.
Furthermore, if [[ ` then
_
_
(1 `)
1
_
_

1
[[ |`|
. (-)
28-1
28-2 28. THE SPECTRUM OF A LINEAR MAP
Proof. It follows from the previous theorem that the resolvent set (`) is open. Also
following the proof of the previous theorem we see that if |`| then
(1 `)
1
=

n=0

(n+1)
`
n
.
so (`). The estimate (-) follows by summing the r.h.s. of
_
_
(1 `)
1
_
_

n=0
[[
(n+1)
|`|
n
.

Theorem 28.5. For any (`) we have


_
_
(1 `)
1
_
_

1
dist(. (`))
.
Proof. For [.[ < 1, |(1 `)
1
| the geometric series
(( +.)1 `)
1
=

n=0
(.)
n
(1 `)
1n
is absolutely convergent, so +. (`).
It follows that for any r A and / A
t
the function
)
x,
() = (1 `)
1
r. /).
has a convergent power series expansion at each point of (`). Thus )
x,
is analytic in (`).
From (-) we have
[)
x,
()[ |r| |/|
1
[[ |`|
for large [[. Thus )
x,
vanishes at innity. By Liouvilles theorem a bounded entire function
is constant. Thus, we have either )
x,
0 or (`) ,= C.
Theorem 28.6. (`) is a non-empty closed subset of the disk : [[ |`|.
Proof. It follows from the previous proof that (`) is closed and contained in [[
|`|. Suppose (`) is empty, so (`) = C. It follows that )
x,
0 for all r A and
/ A
t
. This however is a contradiction as then (1 `)
1
= 0, so 0 is an invertible
operator.
Theorem 28.7. (`) = (`
t
).
Since (1 `)
t
= 1 `
t
, this follows easily from
Lemma 28.8. 1 (L(A) if and only if 1
t
(L(A
t
).
Proof. If 1 is invertible then 11 = 11 = 1
X
with 1 = 1
1
. By taking adjoints
1
t
1
t
= 1
t
1
t
= 1
X
.
so 1
t
1
= [1
1
]
t
.
Supposes now that 1
t
is invertible. Then ran 1
t
= A
t
, so the null space of 1 (which
is the annihlator of ran 1
t
) is 0. Thus 1 is one-to-one. To see that 1 is onto, note that
since `
K
= 0 it follows that ran 1 is dense. Thus it suces to show ran 1 is closed. To
28. THE SPECTRUM OF A LINEAR MAP 28-3
prove this, note that 1
tt
(L(A
tt
) by the rst part of the theorem and that 1
tt
r = 1r
for r A. However, A is a norm closed subspace of its double dual, since the norm on
the double dual is the same as the norm on A (by the dual characterization of the norm
|r| = sup

[r. /)[). Applying the following lemma with 1 = A


tt
and 1 = [1
tt
]
1
it follows
that 1 has closed range.
Lemma 28.9. Let 1 L(A. 1 ) and suppose there is 1 L(1. A) such that 11 = 1
X
.
Then `
X
= 0 and ran 1 is closed.
Proof. Clearly `
X
= 0. On the other hand if 1r
n
is a convergent sequence in
ran 1, then r
n
1 so 1r
n
11 = and ran 1.

In nite dimensions the spectrum of a linear operator is the same as the set of eigenvalues.
This follows since if A is nite dimensional then 1 L(A) is invertible if and only if it is
one-to-one `
X
= 0. In innite dimension, an operator 1 may fail to be invertible for a
number of reasons:
(1) `
K
may be non-trivial.
(2) ran
K
is contained in a proper closed subspace of A, in which case `
K
is non-trivial.
(3) ran
K
is a proper dense subspace of A
It follows that (`) must satisfy (at least) one of the following:
(1) is an eigenvalue. That is there is non-zero r such that `r = r.
(2) The range of 1 ` is contained in a proper closed subspace of A, in which case
is an eigenvalue of `
t
.
(3) The range of 1 ` may be a proper dense subspace of A.
All three possibilities occur, so spectrum and spectral theory in innite dimensions
is a good deal more complicated than in nite dimensions.
LECTURE 29
Some examples
Reading: 20.2-20.3
Shifts
Let 1 and 1 denote the right and left shifts on c
0
, sequences that vanish at ,
1(c
1
. c
2
. . . .) = (0. c
1
. c
2
. . . .)
1(c
1
. c
2
. . . .) = (c
2
. c
3
. . . .).
Let 1
p
, 1
p
denote the restriction of these operators to /
p
, 1 j < and let 1

, 1

denote
the natural extension of these to /

. Note that we have


1
t
= 1
1
and 1
t
= 1
1
.
1
t
p
= 1
p
and 1
t
p
= 1
p
.
1,j + 1,j
t
= 1, 1 j < .
Proposition 29.1. (1) = (1) = (1
p
) = (1
p
) = [[ 1, for each 1 j .
Proof. First note that |1| = |1| = |1
p
| = |1
p
| = 1, so all of the various spectra are
contained in the unit disk. On the other hand for each [[ < 1 the sequence
(.
2
.
3
. . . .)
is an eigenvector of the left shift with eigenvalue . Since this vector lies in
p
/
p
= /
1
and
the spectrum is closed we have (1) = (1
p
) = [[ 1. Since 1
1
= 1
t
and 1
p
= 1
t
p
,
1 < j , it follows that (1
p
) = [[ 1 for 1 j . Likewise, since 1
1
= 1
t
it
follows that (1) = (1
1
) = [[ 1.
Now consider the space c
o
(Z) of two sided sequences. We may dene right and left shifts
here as well

1(. . . . c
1
. c
0
. c
1
. . . .) = (. . . . c
2
. c
1
. c
0
. . . .)

1(. . . . c
1
. c
0
. c
1
. . . .) = (. . . . c
0
. c
1
. c
2
. . . .).
Note that these operators are isometries and inverses of each other

1

1 =

1

1 = 1.
Proposition 29.2. (

1) = (

1) = [[ = 1.
We use
Lemma 29.3. Let ` L(A) be an isometry onto A. Then (`) [[ = 1.
29-1
29-2 29. SOME EXAMPLES
Proof. Since |`r| = |r| for all r, ` is 1 1. As ` is also onto, 0 (`). As
|`| = 1, it follows that (`) [[ 1 and that [[ < 1 (`), as the geometric
series
( `)
1
=

n=0

n
`
n1
converges there.
Proof of Proposition. It follows from the lemma that the two spectra are subsets
of the circle. We show that ran(

1) ,= c
0
(Z) for each of modulus one. The proof for

1
is similar.
Suppose (1

1)a = b. Then the coecients satisfy
c
j
c
j1
= /
j
.
Since c
j1

1
c
j2
=
1
/
j1
it follows that
c
j

1
c
j2
= (/
j
+
1
/
j1
).
Continuing we conclude that
c
j

n
c
jn1
=
n

m=0

m
/
jm
.
Since c
jn1
0 as : we nd that
c
j
= lim
n
n

m=0

m1
/
jm
.
Thus, for instance, the sequence /
j
=
j 1
[j[+1
cannot be in ran(1

1) since
n

m=0

m1
/
jm
=
j1
n

m=0
1
[, :[ + 1
diverges as : .
Note an amusing point of the proof: we actually derived a formula for the inverse of
(1

1). The point is that this inverse is dened only on a dense domain in A.
Similarly, we have
Proposition 29.4. Let

1
p
,

1
p
be the right and left shifts on /
p
(Z) for 1 j . Then
(

1
p
) = (

1
p
) = [[ = 1.
The case j = is easy since then every point of the circle is an eigenvalue of each shift.
The remaining cases are left as an exercise.
Volterra Integral Operators
Consider the operator of integration
\ )(r) =
_
x
0
)()d
on the Banach space A = C[0. 1].
VOLTERRA INTEGRAL OPERATORS 29-3
Theorem 29.5. \ is a bounded operator, |\ | 1, and (\ ) = 0. 0 is not an
eigenvalue of \ .
Proof. It is easy to verify that |\ | 1. Clearly \ ) = 0 if and only if ) = 0 so 0 is not
an eigenvalue of \ . Laxs proof that (\ ) = 0 relies on results from the theory of Banach
algebras which we have not proved yet, so let us proceed directly. We will give a formula for
the inverse of 1 \ . Note that
\
n
)(r) =
_
x
0
_
y
1
0
. . .
_
y
n1
0
)(
n
)d
n
d
n1
d
1
.
It follows that
|\
n
)| |)|
_
1
0
_
y
1
0
. . .
_
y
n1
0
)(
n
)d
n
d
n1
d
n
=
1
:!
|)| .
Thus the geometric series

n=0
1

n+1
\
n
is norm convergent with norm bounded by [[
1
e
[[
1
. Clearly
(1 \ )

n=0
1

n+1
\
n
= 1.
Thus 1\ is boundedly invertible for all ,= 0. Since the spectrum is non-empty we must
have (\ ) = 0.
One can also verify directly that \ is not invertible. After all \ )(0) = 0 so ran \ is
contained in the space of functions that vanish at 0. Note that \
t
is the map on /[0. 1]
which maps a measure dj to the measure j(r. 1]dr. That is \
t
(j)(o) =
_
S
j((r. 1])dr for
any measurable set. Thus 0 is an eigenvalue of \
t
with eigenvector
0
dr.
Part 8
Compact Linear Maps
LECTURE 30
Compact Maps
Motivation: integral operators
Let 1 be a compact Hausdor space and j a Borel measure on 1. Suppose 1(:. t) is a
continuous function on 1 1 and dene a map K L(C(1)) by
K)(:) =
_
T
1(:. t))(t)dj(t).
Since 1 is continuous on a compact space, hence uniformly continuous and bounded, the
image K) is continuous and satises
max
s
[K)(:)[ const. max
t
[)(t)[.
Thus K is indeed a bounded map from C(1) into itself.
Such maps are certainly on of the most important examples of bounded linear maps.
(There are 1
1
, 1
2
, etc. analogues.) The map is not typical, however, because it has one
additional important property. Suppose [)(t)[ 1 for all t. Then
[K)(:) K)(:)[ [j[(1) sup
t
[1(:. t) 1(:. t)[.
Since the right hand side is nite and converges to zero as : : we nd that
Lemma 30.1. The collection K) : |)|
C(T)
1 is equicontinuous. That is, given
: 1 and c 0 there is a neighborhood l of : such that [K)(:) K)(:)[ c whenever
: l and |)|
C(T)
1.
Since the collection is also uniformly bounded, by the Arzela-Ascoli theorem
Theorem 30.2. The operator K maps the unit ball in C(1) to a pre-compact set.
Thus integral operators as above have the rather special property of mapping the unit
ball of the Banach space onto a small subset of the Banach space. This has many useful
and important consequences.
Compact operators
Definition 30.1. A map C L(A. 1 ) is compact if the image C1
1
(0) of the unit ball
in A is pre-compact in 1 (in the norm topology).
Remark. Recall that a subset o of a complete metric space is pre-compact if its closure
is compact. Equivalently, every sequence in o has a Cauchy subsequence.
Proposition 30.3. If C
1
and C
2
are pre-compact subsets of a Banach space then
(1) C
1
+C
2
is pre-compact.
(2) `C
1
is pre-compact for any ` L(A. 1 ) with 1 a Banach space.
(3) the convex hull of C
1
is pre-compact.
30-1
30-2 30. COMPACT MAPS
Let ((A. 1 ) = compact maps from A 1 .
Theorem 30.4. ((A. 1 ) is closed sub-space of L(A. 1 ). Furthermore if M L(1. 2)
and N L(\. A) then MCN ((\. 2) for any C ((A. 1 ).
Proof. The proof that ((A. 1 ) is a subspace is left as an exercise, based on the propo-
sition above, as is the proof that MC is compact if C is. To see that CN is compact,
note that CN1
1
C1
|N|
, so
1
|N|
CN1
1
is pre-compact and hence so is CN (by (2) of the
proposition).
Finally, suppose C
n
C in L(A. 1 ). Let r
j
be a sequence in the unit ball of A. Let
r
j;1
be a subsequence such that C
1
r
j;1
converges as , . By induction construct r
j;n
for
each : such that
(1) r
j;n
is a subsequence of r
j;n1
(2) lim
j
C
n
r
j;n
=
n
.
It follows that

m
= lim
j
C
m
r
j;n
for any : :. Note that
m

n
= lim
j
(C
m
C
n
)r
j;n
, so
|
m

n
| |C
m
C
n
| .
Thus
n
are Cauchy and have a limit . For each :, let ,
n
be such that
|
n
C
n
r
jn;n
| |C
n
C| .
Consider the diagonal sequence r
jn;n
. Then
|Cr
jn;n
| |CC
n
| +|C
n
r
jn;n

n
| +|
n
| 3 |C
n
C| 0.
Thus Cr
jn;n
is a convergent subsequence of Cr
j
and so C1
1
(0) is compact.
Let ((A) = ((A. A).
Theorem 30.5. Let C ((A), let I be the identity map on A, and let T = I C. Then
(1) `
T
is nite dimensional.
(2) There is an integer i such that
`
T
k = `
T
i for / i.
(3) ran T is closed.
Proof. Note that r `
T
i r = Tr. Thus the unit ball of `
T
is contained in T1
1
(0).
Thus the closed unit ball of `
T
is compact. So `
T
is nite dimensional. (See Lecture 5.)
Note that `
T
k+1 `
T
k, and that each subspace is nite dimensional (since T
k
is compact
for any /). Suppose this sequence never stabilizes. By Lemma 5.2 there is then a vector r
k
for each / such that
r
k
`
T
k . |r
k
| = 1 and |r
k
r|
1
2
for all r `
T
k1.
Then, if : < :,
C(r
n
r
m
) = r
n
Tr
n
r
m
+Tr
m
.
Now Tr
n
. r
m
. Tr
m
`
T
n1 (since : < :), so
|Cr
n
Cr
m
|
1
2
.
COMPACT OPERATORS 30-3
So Cr
n
has no Cauchy subsequence, which is a contradiction.
To prove ran T is closed, let
k
= Tr
k
and suppose
k
. We may shift r
k
as we like
by elements of the `
T
. So r
k
may get very large. However, if we consider the distance
d
k
= inf
uN
T
|r
k
n| .
from r
k
to the `
T
, then I claim we have
sup
k
d
k
< .
Indeed, suppose not. Then, passing to a subsequence so d
k
and choosing n
k
so
|r
k
n
k
| 2d
k
we have
0 = lim
k

k
d
k
= lim
k
T
r
k
n
k
d
k
.
since
k
are bounded. As |r
k
n
k
| ,d
k
2 we may pass again to a subsequence to conclude
that
C
r
k
n
k
d
k
converges. Thus
lim
k
r
k
n
k
d
k
= . = lim
k
C
r
k
n
k
d
k
= C..
so T. = 0, that is . `
T
. But then |r
k
n
k
d
k
.| d
k
so
_
_
_
x
k
u
k
d
k
.
_
_
_ 1 contradicting
the convergence derived above. Thus sup
k
d
k
< .
Now since d
k
are bounded, we may choose n
k
`
T
so that |r
k
n
k
| 2 sup
k
d
k
. Now
pass to a subsequence so that
C(r
k
n
k
)
converges. But then
lim
k
(r
k
n
k
) = + lim
k
C(r
k
n
k
).
Thus lim
k
(r
k
n
k
) = r exists and = Tr ran T.
LECTURE 31
Fredholm alternative
Reading: 21.1
Theorem 31.1 (Fredholm Alternative). Let C ((A). Then T = I C satises
dim`
T
= dim(A, ran T).
Remark. The result is known as the Fredholm Alternative because Fredholm proved
this result in the context of integral operators. The alternative is that either one is an
eigenvalue of C, or the equation
= r Cr
is uniquely solvable for r for any .
Corollary 31.2. Let C ((A). Then a non-zero point C is in the spectrum of C
if and only if is an eigenvalue of nite algebraic and geometric multiplicity, where
geometric multiplicity = dim`
IC
and
algebraic multiplicity = sup
j
dim`
(IC)
j .
Proof. Apply the previous two theorems to the compact map
1
C.
Caution: a compact map need not have eigenvalues. The Volterra operator of the last
lecture is an example.
Recall that dimA,1 is the codimension of 1 , denoted The identity
dim`
M
= codimran `
is valid for any operator ` L(A) if A is nite dimensional. In an innite dimensional
setting, this identity need not hold even if both sides are nite. If both sides are nite, the
index of ` is the dierence
ind T = dim`
M
codimran `.
Thus the alternative says that the index of a I C is zero for C compact.
To prove the theorem we will use the following
Lemma 31.3. Let C ((A) and suppose 1 A is an invariant subspace for C, so
C1 1 . Then

C[r] = [cr]
is a compact map from A,1 A,1 .
Proof. Exercise.
31-1
31-2 31. FREDHOLM ALTERNATIVE
Proof of Theorem. First suppose dim`
T
= 0. We need to show that ran T = A.
If, on the contrary, A
1
= ran T is a proper subspace of A. Since T is one-to-one, A
2
= TA
1
is a proper subspace of A
1
. By induction, with A
k
= T
k
A, we nd that A
1
A
2
A
3

with proper inclusion at every step. Now A
1
is closed by Thm 30.5 from last time. Likewise
A
k
= T
k
A and T
k
= I

k
j=1
_
k
j
_
(1)
j1
C
j
= I compact, so A
k
is closed. Thus, we may
nd a sequence of vectors r
j
such that
r
j
A
j
. |r| = 1 . and dist(r
j
. A
j+1
)
1
2
.
It follows, if : < :, that
|Cr
m
Cr
n
| = |x
m
Tr
m
r
n
+Tr
n
|
1
2
.
since Tr
m
+ r
n
Tr
n
A
m+1
. Thus Cr
n
has no Cauchy subsequence, which is a contra-
diction.
If dim`
T
0, then by Thm 30.5 from last time, we have `
T
i+1 = `
T
i for i large enough.
Let ` = `
T
i . So ` is an invariant subspace for T, and since C = I T we see that `
is invariant for C as well. Thus

C as dened above is a compact map of A,` A,`.
Consider

T =

I

C. That, is

T[r] = [1r].
which is well-dened since ` is invariant under 1. Note that `
e
T
= 0, since if [r] `
e
T
then 1r ` = r `
T
i+1 = `
T
i = ` so [r] = 0. Thus by ran

T = A,`. Thus for
every A there is r A and . ` such that
Tr = +..
Thus A = ran T+`. Thus
dim(A, ran T) = dim(`,(` ran T)).
By nite dimensional linear algebra
dim(`,(` ran T)) = dim`
T
.
Theorem 31.4 (Schauder). C L(A) is compact if and only if C
t
is compact.
Proof. We show C compact = C
t
is compact. The reverse implication follows by
noting that C is the restriction of C
tt
to the closed subspace A of A
tt
and applying the
following
Lemma 31.5. Let C be compact on A and let 1 A be a closed subspace. Then the
restriction of C to 1 is a compact map.
Proof. Exerise.
We must show, if /
n
1
1
(0) A
t
, that C
t
/
n
has a convergent subsequence. Consider
the closure of 1 = C1
1
(0) A. This is a compact set and on this set
[/
n
(r) /
n
()[ |r |
since |/
n
| 1. Thus /
n
[
K
are equicontinuous, and bounded, so have a convergent subse-
quence by Arzela-Ascoli. That is, passing to a subsequence, we have
lim
n
sup
yK
[. /
n
)()[ = 0
31. FREDHOLM ALTERNATIVE 31-3
for some cts function ) on 1. Thus
sup
xB
1
(0)
[r. C
t
/
n
)(Cr)[ 0.
It follows that /(r) = )(Cr) is a linear functional and the norm limit of C
t
/
n
. So C
t
is
compact.
Theorem 31.6 (Fredholm Alternative). Let C ((A), T = I C. Then
(1) r ran T if and only if /(r) = 0 for all / `
T
.
(2) dim`
T
= dim`
T
.
Proof. (1) For a general operator T we have
ran T = r : /(r) = 0 fo all / `
T
.
Since ran T is closed (1) follows.
(2) The null space `
T
is isomorphic to the dual of A, ran T via the pairing
[r]. /) = r. /)
which is well dened since / annihlates the ran T. Thus
dim`
T
= codimran T = dim`
T
.
LECTURE 32
Spectral Theory of Compact Maps
Theorem 32.1 (F. Riesz). Let A be a complex Banach space and C ((A). The
spectrum of C is a denumerable or nite set of points whose only accumulation point, if any,
is 0. If dimA = then 0 (C):
(C) =
j
0.
Furthermore,
(1) Each non-zero
j
(C) is an eigenvalue of nite algebraic and geometric multi-
plicity.
(2) The resolvent (. C)
1
has a pole at each non-zero
j
: that is, there is :
j
1 such
that
(.
j
)
n
j
(. C)
1
r. /)
is analytic in a neighborhood of
j
for every / A
t
and r A.
Proof. We have already seen that any non-zero point of the spectrum is an eigenvalue
of nite multiplicity.
Suppose we have a sequence
n
of eigenvalues and eigenvectors
Cr
n
=
n
r
n
.
with
n
,=
m
. Let 1
n
be the linear space of r
1
. . . . . r
n
. Suppose
n

j=1
c
j
r
j
= 0.
Then
n

j=1
c
j

k
j
r
j
= 0
for all /, so
n

j=1
c
j
j(
j
)r
j
= 0
for all polynomials j. Since the
j
are distinct, we may pick a polynomial that vanishes for
all
j
except
j
0
for example j() =

j,=j
0
(
j
). Thus c
j
= 0 for all ,, so r
j
are linearly
independent. It follows that 1
n1
is a proper subspace of 1
n
.
The implication of this is the same as above: we nd a sequence of vectors
n
1
n
such
that
|
n
| = 1 and |
n
|
1
2
for all 1
n1
.
Since

n
=
n

j=1
c
(n)
j
r
j
32-1
32-2 32. SPECTRAL THEORY OF COMPACT MAPS
we have
C
n
=
n

j=1
c
(n)
j

j
r
j
so
C
n

n
r
n
=
n1

j=1
c
(n)
j
(
j

n
)r
j
1
n1
.
Thus, for : :,
C
n
C
m
=
n

n
for some 1
n1
.
Thus
|C
n
C
m
|
[
n
[
2
.
Since any subsequence of C
n
has a Cauchy subsequence, we must have
n
0. Since
n
was an arbitrary sequence of eigenvalues, it follows that 0 is the only (possible) accumulation
point of eigenvalues. Thus there are only nitely many eigenvalues outside any disk around
the origin. Hence there at most countably many eigenvalues.
It remains to show that the resolvent has poles at
j
. Let . be a complex number, then
nding the resolvent n = (. C)
1
r amounts to solving
r = .n Cn
for r. Suppose . is close to
j
. Let ` = `
(C)i
= `
(C)
i+1 with i suciently large. Let

C
be the quotient map on A,`. Since
j
is not an eigenvalue of

C (why?), we conclude that

j


C is invertible. It follows that .

C is invertible for .
j
< c for some c 0 (recall
that the set of invertible maps is open) and
_
_
(. C)
1
_
_
const. for [.
j
[ < c.
Thus
.[]

C[] = [r]
has a unique solution for an equivalence class [] = +` with
|[]| const. |[r]| .
Thus given r we may nd : ` and (.) A such that
(. C)(.) = r :(.).
We may choose (.) so that |(.)| is bounded for [.
j
[ < c. It follows that |:(.)| is
bounded as well:
|:(.)| ([.[ +|C|) |(.)| +|r| .
Let us now solve
(. C)(.) = :(.)
for (.) ` to obtain n(.) = (.) + (.). This is a linear algebra problem. We know that
` is an invariant subspace for C and that
j
is in the spectrum (C [
N
). In fact,
j
is the
unique point of the spectrum of C [
N
. Indeed, (
j
C)
i
= 0 on `. It follows that
_
_
(. C [
N
)
1
_
_
const.
1
[ .[
i
.
(It doesnt really matter for the proof, but yes it is the same i in both spots.) Hence,
|n(.)| |(.)| +|(.)| const.(1 +[.
j
[
i
)
32. SPECTRAL THEORY OF COMPACT MAPS 32-3
for . close to
j
.
Thus (.
j
)
i
(.C)
1
r. /) is analytic in 0 < [.
j
[ < c with a removable singularity
at . =
j
. That is (. C)
1
r. /) has a pole at
j

Note that the resolvent may not have a pole at 0. For instance, the Volterra integral
operator has resolvent
( \ )
1
=

n=0
1

n+1
\
n
for ,= 0. This analytic operator valued function has an essential singularity at 0. If one
computes
( \ )
1
r
n
=

m=0
1

m+1
\
m
r
n
=

m=0
1

m+1
:!
(: +:)!
r
n+m
= :!
n1

m=n
1
:!
_
r

_
m
= :!
n1
e
x/

n1

m=0
:!
:!
r
m

n1m
.
then one can see the essential singularity very clearly in the rst term. The second term is
regular and has limit
:r
n1
as 0.
Interestingly, this is \
1
r
n
where \
1
is the densely dened left inverse for \ , namely
dierentiation:

x
\ )(r) =
x
_
x
0
)()d = )(r).
Homework III
Homework III
Due: April 30, 2008
This homework assignment deals with elliptic PDEs. Let R
d
be an open set with
compact closure. Suppose we are given a measurable function (r) on which is bounded
above and below:
0 < c (r) c
1
< for all r .
Consider the PDE
(r)n(r) = )(r) r . n = 0 on . (-)
Our goal is to show that this equation has a unique solution n = o) with o a compact
symmetric operator on 1
2
and to consider the spectral theory of o.
(1) Show for any ) 1
2
() that (-) has a unique solution (in the sense of distributions) n
in 1
2
(). (Hint: use Lax Milgram. The solution is outlined in Lecture 16.)
(2) Let (r) be a smooth, non-negative, compactly supported, function on R
d
with
_
R
d
(r)dr =
1. Let 1
t
map 1
2
() into itself by
1
t
)(r) =

(r)
_
R
d
t
d
(
r
t
))()d =

(r)
_
R
d
t
d
(

t
))(r )d.
where we take ) 0 outside .
(a) Use Minkowskis inequality:
_
_
dj(r)

_
)(r. )d()

2
_1
2

_
d()
__
dj(r) [)(r. )[
2
_1
2
to show that 1
t
L(1
2
() with |1
t
|
L
2
()L
2
()
1.
(b) Show also that
sup
x
[1
t
)(r)[ Ct
d
|)|
L
2
()
.
(Hint: the constant will be proportional to Vol().)
(c) Show that for ) 1
2
() and r
)(r) 1
t
)(r) =
_
R
d
t
d
(

t
)()(r) )(r ))d.
(d) Apply Minkowskis inequality to conclude, if ) C
2
c
() then
|) 1
t
)|
L
2
()
Ct |)|
H
1
0
()
. (--)
where
|)|
2
H
1
0
()
=
d

j=1
|
j
)|
2
L
2
()
.
(Hint: write )(r) )(r ) =
_
1
0
)(r :))d:. Make sure you get the factor
of t on the r.h.s!)
32-2
(e) Conclude that (--) holds for ) H
1
0
() with
H
1
0
() = completion of C

c
() in the norm |)|
2
H
1
=
d

j=1
|
j
)|
2
L
2
.
(3) Let )
n
H
1
0
() be a sequence so that
j
)
n

j
weakly in 1
2
() for each ,.
(a) Use Lemma 16.1 of Lecture 16 (Lemma 2 of Ch. 7 in Lax) to conclude that )
n
)
weakly in 1
2
() for some ) H
1
0
() and that
j
) =
j
. (Hint: rst use Alaogulus
theorem to conclude that )
n
has a weakly convergent subsequence with a limit )
that satises
j
) =
j
for each ,. Next show that any other weakly convergent
subsequence must also converge to ). Conclude that )
n
).)
(b) Show that
1
t
)
n
(r) 1
t
)(r) as : .
for every r . Since [1
t
)
n
(r)[ Ct
d
|)
n
|
L
2
()
by (2b) use dominated convergence
and the principle of uniform boundedness to conclude that
|1
t
)
n
1
t
)|
L
2
0 as : .
(c) Use this result and (--) to conclude that
|)
n
)|
L
2
0 as : .
(d) Combine these to prove
Theorem 32.2 (Rellich). If ` H
1
0
() with uniformly bounded H
1
0
() norm then
` is pre-compact in 1
2
().
(4) Returning to the PDE (-), let o : 1
2
() 1
2
() be the map o) = n where n solves
(-). Show that o is compact. (Hint: see 22.3 where this is done for 1. However, we
have not assumed that has a smooth boundary. But you have just proved the results
needed.)
(5) Show that o is symmetric on 1
2
(). Conclude that there is an ortho-normal basis
j

for 1
2
() consisting of eigenfunctions to the linear operator (r). That is, there is
a set
j
such that
span
j
= 1
2
() .
j
.
k
)
L
2
()
=
i,j
.
and
(r)
j
(r) =
j

j
(r).
Show that the eigenvalues
j
are real and diverge to . How are they related to the
eigenvalues of o?
Part 9
Compact Linear Maps in Hilbert Space
LECTURE 33
Compact Symmetric Operators
We now specialize to the case in which the Banach space A is a Hilbert space. For
spectral analysis it is useful to consider a complex Hilbert space H, in which case in place
of the transpose it is natural to consider
Definition 33.1. The adjoint of a bounded linear map 1 L(H. 1) with H and 1
Hilbert spaces is the linear map 1

L(1. H) given by
1

n. )
H
= n. 1)
K
for all n 1 and H.
Recall that, by the Riesz-Frechet theorem the dual of a Hilbert space H is isomorphic to
the Hilbert space itself via the conjugate linear map
n /
u
() = . n).
Thus we have
1
t
/
u
= /
T

u
.
Since the correspondence n /
u
is conjugate linear, it follows that
(1 +co)

= 1 +c

.
We also have
(o1)

= 1

and
(1

) = (1)

: (1).
Definition 33.2. A bounded operator 1 L(H) is called Hermitian (or symmetric, or
self-adjoint) if 1 = 1

.
Proposition 33.1. If 1 is Hermitian then
(1) 1n. n) is real for all n H.
(2) If 1n. n) = 0 for all n H then 1 = 0.
(3) (1) R.
Proof. (1) Note that 1n. n) = n. 1n) = 1n. n)

. For (2), note that if 1n. n) = 0 for


all n then
0 = 1(n i). n i) = i1. n) i1n. ).
Combining the results we get
0 = 1n. )
for all n. H. Thus 1 0.
For (3), we need to show `
zuT
= 0 and ran(. 1) = H, for . C R. Suppose
.n 1n = 0.
33-1
33-2 33. COMPACT SYMMETRIC OPERATORS
It follows that
. |n|
2
1n. n) = 0.
Since the second term is real we nd Im. |n|
2
= 0. Thus |n| = 0 and n = 0. That is,
`
zuT
= 0.
Similarly, suppose n is perpendicular to ran(. 1). Then
0 = (. 1)n. n).
so |n| = 0. It follows that ran(. 1) is dense. To see that ran(. 1) is closed, note
|(. 1)n|
2
= (. 1)n. (. 1)n)
= [.[
2
|n|
2
+|1n|
2
.n. 1n) .

1n. n)
= [.[
2
|n|
2
+|1n|
2
2 Re .1n. n)
= (Im.)
2
|n|
2
+|(Re . 1)n|
2
(Im.)
2
|n|
2
.
Thus, if ran(. 1)n
n
r we nd that n
n
is Cauchy. since (. 1) is bounded we must
have (. 1)n = r for n = lim
n
n
n
.
Theorem 33.2. Let H be a innite dimensional separable Hilbert space. If 1 L(H) is
Hermitian and compact then there is an orthonormal basis for H consisting of eigenvectors
of 1. That is, there is an orthonormal basis
n

n=1
and a sequence of real numbers
n
0
such that
1
n
=
n

n
.
The spectrum (1) =
n

n=1
.
Remark. If H is not-separable and 1 is compact Hermitian then there is a closed
separable subspace 1 H such that 1 : 1 1 and
1[
Y
= 0.
Note that we already have (H) =
n
R. To proceed let us show rst
Lemma 33.3. Let 1 ((H). Let
+
= max (1) and

= inf (1). Then

+
= sup
|u|=1
1n. n) = max
|u|=1
1n. n)
and

= inf
|u|=1
1n. n) = min
|u|=1
1n. n).
In particular, if (1) = 0 then 1 0.
Proof. Let us prove the identity for
+
. The identity for

follows from the rst


applied to 1. Let 1 = sup
|u|=1
1n. n). Note that
1n. n) 1|n|
2
.
It follows that 1 0 as otherwise [1n. n)[ [1[ |n|
2
. so [1[ |n| |1n| . which contradicts
compactness. (Look at 1n
n
with n
n
an orthonormal basis.)
Let n
n
be a sequence of vectors with |n
n
| 1 so that 1n
n
. n
n
) 1. Passing to a
subsequence, we suppose that n
n
n
+
and 1n
n
1n
+
(strongly). It follows that
1n
n
. n
n
) 1n
+
. n
+
).
33. COMPACT SYMMETRIC OPERATORS 33-3
so 1n
+
. n
+
) = 1. Thus the sup is actually a max.
If 1 then, for |n| 1,
|( 1)n| 1n. n)
2
1.
It follows that
|( 1)n| ( 1) |n|
2
for all n. Thus 1 0 is not an eigenvalue of 1. It follows that
+
1.
If 1 = 0, we are done as
+
0. On the other hand, if 1 0 then n
+
,= 0, since
1 = 1n
+
. n
+
. Clearly n
+
is a unit vector. Let u H and consider the function
1
w
(t) =
1(n
+
+tu). n
+
+tu)
|n
+
+tu|
2
.
So 1
w
is smooth and takes its maximum at 0. Dierentiating with respect to t we get
0 = 1
t
w
(0) = 1u. n
+
) +1n
+
. u) 1n
+
. n
+
) [u. n
+
) +n
+
. u)] .
Thus
0 = 1
t
w
(0) + i1
t
iw
(0) = 1n
+
. u) 1n
+
. u).
Since this holds for any u, we conclude that 1n
+
= 1n
+
, so 1 (1).
Finally, note if (1) = 0 then 1n. n) = 0 for all n so 1 0 by the proposition proved
above.
Proof of Theorem. If 1 0 there is nothing to show as any orthonormal basis will
do. If 1 is not zero, then it has a non-zero eigenvalue
1
by the lemma. Let the corresponding
eigenvector be n
1
. If n n
1
then
1n. n
1
) =
1
n. n
1
) = 0.
so 1n n
1
. Thus n
1

is an invariant subspace for 1. Either 1 is identically zero on n


1

or it has an eigenvector there. Proceed by induction.


More formally, consider the collection o of closed subspaces 1 of H invariant under 1
and such that 1 restricted to 1 has an orthonormal basis of eigenvectors. As above 1

is
invariant under 1 so either 1 0 on 1

or 1 has a non-zero eigenvector there. Partially


order o by inclusion. It follows that any maximal element 1 of o has 1

= 0.
LECTURE 34
Min-Max
Let us look at the variational characterization of eigenvalues again. We have

+
= max (1) = max
|u|1
1n. n).
and a similar identity at the bottom of the spectrum. As it turns out, one can construct
in this way all of the eigenvalues. Indeed, suppose we are given the largest ` 1 positive
eigenvalues
1

2

N1
0 of 1 (listed according to multiplicity) together with
the corresponding eigenvectors r
1
. . r
N1
. Let

N
= max
ux
1
,...x
N1
, |u|1
1n. n). (-)
If
N
0 then
N
is an eigenvalue of 1,
N

N1
, and the corresponding maximizer n
N
is an eigenvector. If
N
= 0 then there are no more positive eigenvalues. (Zero may or may
not be an eigenvalue.) Similarly we can nd the negative eigenvalues and corresponding
eigenvectors by minimizing 1n. n).
Theorem 34.1 (Min-Max Principle). Let 1 be a compact Hermitian operator on an
innite dimensional Hilbert space.
(1) Fischers principle: If

N
= max
S
N
min
uS
N
, |u|=1
1n. n).
where the max is taken over all ` dimensional subspaces o
N
, then
1

2
0
and the non-zero entries of this sequence are the positive eigenvalues of 1 listed
according to multiplicity.
(2) Courants principle: The above sequence is also given by

N
= min
S
N1
max
uS
n1
. |u|1
1n. n).
where the max is taken over all ` 1 dimensional subspaces o
N1
.
Remark. (1) In nite dimensions, the sequence is no longer non-positive and gives all
the eigenvalues. In innite dimensions, the sequence
j
, being non-positive and decreasing,
is either positive for all , or eventually 0. The zero entries may not represent eigenvalues.
(2) The negative eigenvalues j
1
j
2
< 0 can of course be found as
j
N
= min
S
N
max
uS
N
, |u|=1
1n. n) = max
S
N1
min
uS
N1
, |u|1
1n. n).
Proof. First note that
N
0. Indeed, for any 0 there are only nitely many
eigenvalues < (counted according to multiplicity). Thus the orthogonal complement of
the corresponding eigenvectors is innite dimensional. On this subspace we have 1n. n)
|n|
2
, so
N
.
34-1
34-2 34. MIN-MAX
Let r
n
=
n
r
n
be the positive eigenvalues and eigen-vectors listed according to mul-
tiplicity with
1

2
0. (The sequence may terminate.) If there are ` positive
eigenvalues and ` `, let o
N
be an `-dimensional subspace. By nite dimensional linear
algebra, we can nd a unit vector n o
N
perpendicular to r
1
. . . . . r
M
:
n. r
n
) = 0. : = 1. . . . . `.
By (-) we must have 1n. n) 0 so
N
0. On the other hand
N
0 so
N
= 0.
If there are at least ` positive eigenvalues, then with o
N
= span(r
1
. . . . r
N
) we have
min
uS
N
,|u|=1
1n. n) =
N
.
so
N

N
. On the other hand if o
N
is any ` dimensional subspace then as above we may
nd a unit vector n o
N
which is perpendicular to the rst ` 1 eigenvectors. By (-),
1n. n)
N
. Thus
N

N
.
Turning now to Courants principle, let

N
= min
S
N1
max
uS
N1
, |u|1
1n. n).
Since o

N1
is innite dimensional and 1 is compact it follows that
max
uS
N1
, |u|1
1n. n) 0.
Thus
N
0. Also, note that

N
min
S
N1
min
S
N2
S
N1
max
uS
N2
, |u|1
1n. n) =
N1
.
First suppose there are at least ` 1 positive eigenvalues. It follows from (-) that

N

N
=
N
. since r
1
. . . . . r
N1
span an ` 1 dimensional subspace.
If
N
= 0, it follows that
M
= 0 =
M
for ` `.
On the other hand, suppose there are at least ` positive eigenvalues and let o
N1
be an
arbitrary ` 1 dimensional subspace. I claim we may nd a vector in the subspace spanned
by r
1
. . . . r
N
perpendicular to o
N1
. Indeed, if
j
, , = 1. . . . . ` 1, is an orthonormal basis
for o
N1
then we must solve
N

k=1
r
k
.
j
)c
k
= 0.
for all ,. Since ` 1 < ` the matrix r
k
.
j
) has a non-trivial null space, and the resulting
solution n =

k
c
k
r
k
is perpendicular to o
N1
. But then
1n. n) =
N

k=1
[c
k
[
2

k

N
|n|
2
.
Thus
N

N
=
N
.
The min-max principle is incredibly powerful. It is certainly one of the most important
results in applications of functional analysis. Here is an example of what we can do with it:
Definition 34.1. Let . 1L(H) be Hermitian. We say that 1 if
n. n) 1n. n) n H.
Remark. This is a partial order on the set of Hermitian operators.
34. MIN-MAX 34-3
Theorem 34.2. Let 1. o ((H) be Hermitian compact operators with 1 o. If o has
at least ` positive eigenvalues
+
1

+
N
0 then 1 has at least ` positive eigenvalues
j
+
1
j
+
N
0 and
j
+
N

+
N
.
Similarly, if 1 has at least ` negative eigenvalues j

1
j

N
< 0 then o has at least `
negative eigenvalues

N
< 0 and

N
j

N
.
This theorem is an easy consequence of min-max. It is immediate if o and 1 have the
same eigenvectors, but that is not at all necessary for the relation o 1. For instance
_
1 0
0 0
_

_
2 1
1 1
_
.
Similarly we have
Theorem 34.3. Let 1 ((H) be compact Hermitian. If 1 0 then all eigenvalues of
1 are 0.
In fact, we have
Theorem 34.4. Let 1 L(H) be Hermitian. If 1 0 then (1) [0. ).
Proof. Let 0. Then
|(1 +)n|
2
= |1n|
2
+ 2n. 1n) +
2
|n|
2

2
|n|
2
.
It follows that `
T+
= 0 and that ran(1 +) is closed. Since ran(1 +)

= `
T+
= 0
it follows that ran(1 +) = H. Thus 1 + is one-to-one, onto and bounded. By the inverse
mapping theorem (27.5 in these notes) (1 +)
1
is bounded. Thus , (1).
LECTURE 35
Functional calculus and polar decomposition
The spectral theorem states that a compact Hermitian operator 1 is of the form:
1 =

n=1

n
.
n
)
n
. (-)
Here
n
are the eigenvalues and
n
is an orthonormal basis. The notation indicates that
1 =

n=1

n
.
n
)
n
.
In fact, since
n
0 the sum in (-) is absolutely convergent:
_
_
_
_
_
1
N

n=1

n
.
n
)
_
_
_
_
_
= sup
n>N
[
n
[ 0.
If ) : (1) C is a bounded function we dene
)(1) =

n=1
)(
n
).
n
)
n
.
so
)(1) =

n=1
)(
n
).
n
)
n
.
It is an easy calculation to see that if )(r) = j(r) is a polynomial in r then )(1) dened
this way agrees with plugging 1 into the polynomial. For instance,

n=1

2
n
.
n
)
n
=

n=1

m=1

m
.
m
)
m
.
n
) = 1
2
.
by the pairwise orthogonality of
m
. The map ) )(1) is called the functional calculus for
1 and has the following properties:
Theorem 35.1. Let 1 be a compact Hermitian operator. To every bounded function
) : (1) C we assign a unique operator, )(1), such that
(1) )() 1 = )(1) = I.
(2) )() = = )(1) = 1.
(3) The map ) )(1) is an injective homomorphism of the ring of bounded functions
on (1) into the L(H):
() +p)(1) = )(t) +p(1). ()p)(1) = )(1)p(1). and )(1) 0 i ) 0 on (1).
(4) )(1)

= )

(1).
35-1
35-2 35. FUNCTIONAL CALCULUS AND POLAR DECOMPOSITION
(5) The map is an isometry:
|)|

= sup
(T)
[)()[ = |)(1)| .
This map has the properties
(6) If ) : (1) R then )(1) is Hermitian.
(7) If ) : (1) [0. ) then )(1) 0.
(8) If ) : (1) [.[ = 1 then )(1) is a unitary map, that is an isometry of H onto
H.
(9) If lim
0
)() = )(0) = 0 then )(1) is compact.
Remark. Properties (3-5) show that the map ) )(1) is an isomorphism of C

algebras, something we havent dened yet but that we will see in the fall.
Proof. We already dened the map we will show it is unique in a moment. It is clear
that (1) and (2) hold. It is easy to see that () + p)(1) = )(1) + p(1) and the argument
given for 1
2
above extends to a product )p. Thus (3) holds. To see that (4) holds note that
)(1)

n. ) = n. )(1)) =
_
n.

n
)(
n
)

.
n
)
n
_
=

n
)(
n
)

n.
n
)
n
. ) =
_

n
)(
n
)

n.
n
)
n
. )
_
= )

(1)n. ).
Since
|)(1)n|
2
=

n
[)(
n
)[
2
[n.
n
)[
2
|)|

|n|
2
and
)(1)
n
= )(
n
)
n
.
(5) follows. (6) and (7) are easy calculations. To see (8) note that if [)(r)[
2
= 1 then
|)(1)n|
2
=

n
[n.
n
)[
2
= |n|
2
.
(9) is an easy exercise.
Finally to see that the map is unique, note that (1), (2), and (3) specify the map for
polynomials j(r). Since 1 is compact, (1) is discrete away from 0. Thus a bounded map
on (1) is continuous if and only if it is continuous at 0. Thus (5) and Stone-Weierstrass
imply that the map for polynomials extends uniquely to bounded functions continuous at 0.
In fact the map is uniquely dened on all bounded functions, but we will defer the proof of
this to next term.
Corollary 35.2. If 1 0 and is compact Hermitian then there is a unique positive
square root of 1:

1 0 and

1
2
= 1.
The square root is very useful as it allows us to dene
Definition 35.1. Let 1 be a compact operator on a Hilbert space H. The absolute
value of 1, denoted [1[, is the operator

1.
35. FUNCTIONAL CALCULUS AND POLAR DECOMPOSITION 35-3
The absolute value [1[ is compact (why?) so it has eigenvalues, which are all non-negative
since [1[ 0. The eigenvalues of [1[ are called the singular values of 1. Note that
|1|
2
= . 1

1) = . [1[
2
) = |[1[|
2
.
It follows that the map [1[ 1 dened on the range of [1[ is a linear isometry. This
extends to the closure of ran [1[. Let this map be denoted l and dene l to be zero on
ran [1[

= `
[T[
= `
T
. Note that
1 = l[1[.
Theorem 35.3 (Polar decomposition). Every compact operator 1 on a Hilbert space H
may be factored, as
1 = l
with 0 and l

l [
ran A
= I. The map = [1[ and l is uniquely determined if we specify
l 0 on ran

.
Corollary 35.4. Any compact operator 1 on a Hilbert space has a singular value
decomposition
1 =

n=1
j
n
.
n
)
n
.
with j
n
0, j
n
0 and
n
,
n
(possibly distinct) orthonormal sequences.
Proof. Let j
n
,
n
be the eigenvalues/vectors of [1[. Let
n
= l
n
with 1 = l[1[.
In fact, the functional calculus, square root and polar decomposition extend to non-
compact operators. These are some theorems we will prove next term:
Theorem 35.5. Every positive operator 0 on a Hilbert space has a unique positive
square root.
Theorem 35.6 (Polar Decomposition). Every operator 1 L(H) may be factored as
1 = l
with 0 and l

l [
ran A
= I. The map = [1[ =

1 and l is uniquely determined if


we specify l 0 on ran

= `
T
.
Theorem 35.7 (Spectral Theorem: functional calculus version). To every self adjoint
operator 1 L(H) there is associated a unique injective C

homomorphism from C((1))


L(H) with the properties (1)-(8) above. Conversely, given any compact set R and an
injective C

homomorphism with these properties there is a self adjoint operator 1 so that


the homomorphism is realized as ) )(1).

Anda mungkin juga menyukai