Functional Analysisf15
Functional Analysisf15
Yen Do
Fall 2015
2
Preface
This is the accompanying expository notes for an introductory course in Functional Analysis
that I was teaching at UVA. The goal of the course is to study the basic principles of linear
analysis, including the spectral theory of compact and self-adjoint operators. This is not a
monograph or a treatise and of course no originality is claimed. The prerequisite
is some basic knowledge about real analysis and topology. Some preliminary understanding
of functional analysis is beneficial but not required.
3
4
Contents
I Linear Spaces 9
1 Basic facts 11
1.1 Metric spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.1.1 Separability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.2 Linear spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.2.1 The Hahn-Banach extension theorem . . . . . . . . . . . . . . . . . . 15
1.2.2 The HB theorem with symmetry constraints . . . . . . . . . . . . . . 17
1.3 Topological spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
1.3.1 Compactness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
1.3.2 Continuous functions on LCH spaces . . . . . . . . . . . . . . . . . . 19
1.3.3 Proof of Urysohn’s lemma . . . . . . . . . . . . . . . . . . . . . . . . 20
1.3.4 Proof of Tietze’s theorem . . . . . . . . . . . . . . . . . . . . . . . . 21
1.3.5 Partition of unity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
5
6 CONTENTS
9 Compact operators 73
9.1 Compact operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
9.2 Compactness of integral operators . . . . . . . . . . . . . . . . . . . . . . . . 75
9.3 Spectral properties of compact operators . . . . . . . . . . . . . . . . . . . . 76
9.3.1 Riesz’s theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
9.3.2 Spectral properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
CONTENTS 7
12 Fredholm determinant 89
12.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
12.2 Fredholm’s approach for integral operators . . . . . . . . . . . . . . . . . . . 90
12.2.1 Convergence and Continuity of the determinant . . . . . . . . . . . . 91
12.3 Fredholm determinant for operators on Hilbert spaces . . . . . . . . . . . . . 95
8 CONTENTS
Part I
Linear Spaces
9
Chapter 1
Preliminaries
11
12 CHAPTER 1. BASIC FACTS
Theorem 1. If (M, d) is a metric space then there exists a complete metric space (M
f, d)
e
and a map h : M → M
f such that
d(m1 , m2 ) = d(h(m
e 1 ), h(m2 ))
h(x) = [(x, x, . . . )]
1.1.1 Separability
We recall that a topological space is separable if there exists a countable dense set. We say
that this set separates the space; for instance the rationals separate the real numbers.
A metric space is said to be separable if the topology induced from the metric is separable.
Note that there is also the concept of second countable, which says that one could find a
countable collection open sets from which one could get all open sets in the topology simply
by taking union.
This would imply separability (just take one point from each such open set), however for
metric/metrizable spaces these two are equivalent.
Separability is important in constructive proofs, for instance we could typically avoid
Zorn’s lemma/the axiom of choice if separability is available.
Basic properties:
1. All compact metric spaces are separable.
2. If the given topological space is an union of a countable number of separable subspaces,
then it is also separable.
Combining Properties 1 and 2, it is clear that an Euclidean space RI (consisting of
functions from I to R) is separable if any only the dimension |I| is countable.
3. A metric space is not separable if there is an uncountable collection of functions such
that the distance between any two is at least 1. (Indeed, if one could find a countable dense
set then at least an element of this dense set has to be close to two functions in the given
collection, which by the triangle inequality would imply that the distance between the two
has to be small, contradictory to the given hypothesis).
Using Property 3, we could show that L∞ [0, 1] (or L∞ (R) etc.) is not separable. To see
this, one could simply take the collection C of characteristic functions of intervals [a, b] ⊂ [0, 1]
where a, b are irrationals – there are uncountably many such functions and any two elements
of C are of L∞ distance at least 1, contradiction.
Using a similar line of reasoning, one could see that the space of all signed Borel measures
on [0, 1] (with norm being the total mass) is not separable. Here simply use the (uncountable)
collection of delta measures at points of [0, 1].
4. We could also define separability for measure spaces by defining a metric on the
underlying σ-algebra. Namely, given a σ-algebra A on X and a corresponding measure µ we
could define a metric on A by
p(A, B) = µ(A∆B)
the measure of the symmetric difference
A∆B = (A ∪ B) \ (A ∩ B)
and we say that (X, A, µ) is separable if the metric space (A, p) is separable.
Lemma 1. If a σ-algebra is generated from a countable collection of sets, then it is separable.
Proof. Let A be a countable generating set, let A0 consist of all finite intersections of elements
of A, and A00 consist of all finite unions of element of A0 , clearly A00 is still countable and it
14 CHAPTER 1. BASIC FACTS
is an algebra of set (i.e. the union and intersection of any two elements are still inside A00 ).
Clearly A00 also generates the given σ-algebra A, thus to show separability it suffices to show
that given any E ∈ A and any > 0 one could fine F ∈ A00 such that µ(E∆F ) < . This is
done by looking at the set of all such E and show that it is actually a σ algebra containing
A00 , thus by minimality it has to contain A and the desired claim follows.
Exercise: Rn with the usual Lebesgue measure is separable.
Theorem 2. Let (X, A, µ) be a separable measure space. Then for any 1 ≤ p < ∞ the space
Lp (X, A, µ) is separable.
Intuitively, the graph of p over X looks like a convex cone with top at the origin.
Note that this is equivalent to
One could think of this theorem as an example of a local to global principle for linear
maps: as long as the local constraint is reasonable we could preserve it globally. Here we
need homogeneity and convexity.
Proof of the HB theorem in this generality uses Zorn’s lemma which is equivalent to the
axiom of choice. (for more concrete X such as Lp , `p , one could avoid the axiom of choice.)
A relation R on a set X is a subset of X ×X (could be empty!). We say xRy if (x, y) ∈ R.
R is an equivalence relation if three conditions holds:
Proof of the HB theorem. We divide the proof into two steps. In step 1, we show that
if Y is a proper subspace of X then we can increase the dimension of Y while preseving the
given hypothesis. In step 2, by induction on the dimension of Y (iterating step 1) and using
Zorn’s lemma we obtain the whole space X.
Step 1: Assume existence of z ∈ X \ Y . Let Ye = span(Y, z) = {ay + bz, a, b ∈ R, y ∈ Y }.
Want to define `0 such that `0 (y) = `(y) and
thus
b`0 (z) ≤ p(ay + bz) − a`(y)
all a, b ∈ R and y ∈ Y . Just need a = 1.
By considering b > 0 and b < 0 we end up needing
1
sup − [p(y − cz) − `(y)]
c>0,y∈Y c
1
≤ inf [p(y + cz) − `(y)]
c>0,y∈Y c
After algebraic manipulations this follows from the given convexity assumption on p,
more specifically the following estimate, c1 , c2 > 0,
c1 c2 c1 c2
p( x+ y) ≤ p(x) + p(y)
c1 + c2 c1 + c2 c1 + c2 c1 + c2
1.2. LINEAR SPACES 17
Step 2: Here we iterate step 1 and use Zorn’s lemma. If X is finite dimensional then it
is clear that the process will stop after finitely many iterations. If X is infinitely dimensional
(the dimension of X could even be uncountable), we need to be able to go beyond any
linearly ordered chain of these extension in order to apply Zorn’s lemma.
Formally, consider the set of all possible extensions of ` from Y to bigger subspaces of
X while remain dominated by p. Define a partial order where (Yα , `α ) ≤ (Yβ , `β ) if Yα ⊂ Yβ
and `α agrees with `β on Yα . Now any linearly ordered chain (Yα , `α )α∈I (here I could be
uncountable) has an upper bound (Y 0 , `0 ) defined by
[
Y0 = Yα
α∈I
all |a| + |b| > 0 complex numbers. Assume that ` : Y → C linear functional such that
|`(y)| ≤ p(y) all y ∈ Y . Then can extend ` linearly to X so that it is still dominated by p(y).
Note: the extension given in the HB theorems is not necessarily unique!
Proof. Idea of proof: `1 (y) = Re(`(y)) and `2 (y) = Im(`(y)) are real functionals bounded
by p, and `2 (y) = −`1 (iy). So
`(y) = `1 (y) − i`1 (iy)
By real HB we could extend `1 to X, thus we extend ` to. But still need to show that
the extended ` satisfies
|`(x)| ≤ p(x)
For any x, let α be such that |α| = 1 and `(x) = α|`(x)|, then the desired estimate
follows:
|`(x)| = `(α−1 x) = `1 (α−1 x) ≤ p(α−1 x) = p(x)
We emphasize again that technically speaking the HB extension theorem applies to any
linear spaces; although heuristically speaking the existence of p implies some implicit ge-
omeric structures.
More specifically, let A be a set of linear maps from X to itself that commute, thus if
T1 , T2 ∈ A then T1 T2 x = T2 T1 x for all x ∈ X. Assume that p is invariant under elements
of A, thus p(x) = p(T x) if T ∈ A and x ∈ X. Assume that (on Y ) ` is invariant under A.
Then
Theorem 4. ` could be extended to all of X so that it remains invariant under A and still
controlled by p.
Idea of proof: We could add more to A all products of its elements, thus we may assume
that A contains 1 and is closed under composition (a semigroup). Let B contain all finite
convex combination of elements of A. By definition, (inside Y ) the given ` is controlled by
p0 (x) = inf T ∈B p(T x), which is a homogeneous convex function on X. Thus we could extend
` to all of X while remains dominated by p0 (thus is dominated by p). We need to show that
` is invariant under A. Let T ∈ A, it can be shown that
p0 (x − T x) ≤ 0
for every x ∈ X. Indeed, by definition p0 (x) ≤ p0 ((1 + · · · + T n−1 )x/n) for any n ≥ 1,
therefore
1
p0 (x − T x) ≤ p(x − T n x) → 0 as n → ∞.
n
Thus `(x − T x) ≤ p0 (x − T x) and thus `(x) ≤ `(T x) for all x. Since `(−x) = −`(x), it
follows that `(x) = `(T x).
An application: The most typical applications of the HB extension theorem are hyper-
plane separation theorems which require some local convexity of the underlying space. We
will revisit these applications later, here we discuss a concrete example.
1.3.1 Compactness
X is compact if for any open covering there is a finite subcover. A space is locally compact
if every point has a compact neighborhood.
Properties: 1. (finite intersection property) For any family of closed set with nonempty
intersection we could find a finite subfamily with nonempty intersection. In fact this is
equivalent to compactness.
2.QTychonoff ’s theorem: If (Xα )α∈A is a family of compact spaces then the product
X = α∈A Xα with the product topology is compact.
(The product topology is the minimal topology such that all coordinate projections πα :
X → Xα are continuous, equivalently speaking it is generated by πα−1 (open sets). )
1.3. TOPOLOGICAL SPACES 19
3. If there is some geometry (i.e. the topology is metrizable) then compactness is equiv-
alent to sequential compactness, which states that for any sequence there is a subsequence
that converges.
4. Net convergence: Without geometry, one needs to use more than sequences. A
net (xα )α∈I is a collection indexed by some directed set I, i.e. a set with some partial
ordering < so that any two elements has at least one common upper bound. We say this net
converges to x if given any neighborhood of x there is β ∈ I such that xα ∈ P if β < α.
(be careful, this limit is not necessary unique in general).
Bolzano–Weierstrass’s theorem: X is compact iff every net has a convergent subnet.
• f = 1 on K and
Tietze’s extension theorem: Let X be LCH and K ⊂ X compact subset. Then any
function f ∈ C(K) could be continuously extended to all of X. Furthermore the extended
function vanishes outside a compact set.
Partition of unity: this ia collection of nonnegative functions whose sum is 1 everywhere
on the space, but locally at each point only finitely many of them are nonzero. One certainly
could only do this partition for a compact subset of the space.
20 CHAPTER 1. BASIC FACTS
Ur ⊂ Us
f (x) > α ⇔ x 6∈ Ur for some r < α. Using the inclusion assumption, this is equivalent
to existence of r > α such that x 6∈ U r , so
[ c
f −1 (α, ∞) = Ur is open.
r>α
Thus the remaining step is to construct the family. Here we use normality: to get W1
open and W2 open that surrounds K and U c (disjoint closed sets). Thus K ⊂ W1 ⊂ W2c ⊂ U ,
and we define U1/2 = W1 , now the closure of U1/2 is inside W2c so inside U . Thus
K ⊂ U1/2 ⊂ U1/2 ⊂ U
1.3. TOPOLOGICAL SPACES 21
We repeat this with the new pairs (K, U1/2 ) and U1/2 , U ) and so on, get the family.
Step 2: Reduction to compact setting. The idea is to show that for some V ⊂ X open
it holds that V is compact and K ⊂ V ⊂ V ⊂ U . Thus if the Lemma holds for compact
Hausdorff, we simply restrict to the subspace V and then extend the local function on V to
the whole of X by letting it be 0 outside V .
The existence of such a V can be done as follows: first we show that given each x ∈ K
we could get a compact neigborhood Nx that remains inside U . Then the family of interior
of these Nx forms an open cover of K, thus using a finite subcovering we easily get a open
set containing K such that its closure is inside U .
Now the existence of such a neighborhood follows from LCH property: the idea is for
Hausdorff space we could actually separate point and a compact set.
Now given each x we could get a neighborhood of x, called Mx , that is compact but not
necessarily inside U , then the set P = Mx ∩ U is compact and is a neighborhood of x, it is
almost contained inside U . One uses Hausdorffness to separate x further from the compact
boundary of this set.
Starting from this chapter, we begin examining linear spaces with at least one extra structure
(topology or geometry). We assume linearity; this is a natural feature of functional spaces.
In this chapter, we start with spaces whose geometric and linear structures are compatible.
More precisely, we assume that the metric is translation invariant d(x, y) = d(x + z, y + z)
and homogeneous d(λx, λy) = |λ|d(x, y). We then define kxk = d(x, 0) the norm of x, and
it is clear that k.k satisfies the following three properties (below x, y ∈ X and λ ∈ C):
• kx + yk ≤ kxk + kyk.
• kλxk = |λ|kxk.
A linear space equiped with such a norm is called a normed linear space (NLS).
Conversely, if such a norm exists we could always define a metric d(x, y) = kx − yk that
is translation invariant and homogeneous.
If the norm topology is separable then we say the NLS is separable.
A Banach space is a complete normed linear space, i.e. the induced metric space is
complete. A Hilbert space has additional structures, where we could talk about “angle”
through the scalar product. We’ll discuss these spaces further in separate chapters.
We recall that an incomplete metric space could be completed, and this applies to normed
linear spaces too: the metric remains invariant under dilation and translation in the com-
pleted space thus it remains a norm.
Some examples:
1. Let (X, µ) be a measure space. Let 1 ≤ p ≤ ∞. Then Lp (X, µ) are normed linear
spaces, with
Z
kf kp = ( |f (x)|p dµ)1/p
X
23
24 CHAPTER 2. OVERVIEW OF NORMED LINEAR SPACES
If X is not compact then the sup is not necessarily finite and so this is not even a norm.
For locally compact (Hausdorff) spaces we could instead look at Cc (X) consisting of
compactly supported continuous functions, and this space is a normed linear space. However,
Cc (X) is not complete in the noncompact case, infact its completion is C0 (X) the space of
continuous functions on X that vanish at ∞. (These are functions f ∈ C(X) such that for
every M > 0 the set {|f | ≥ M } is compact.)
To see this, we first show that C0 (X) is complete under the uniform norm. One way to
show this is to use one-point compactification. Alternatively, consider a Cauchy sequence
(fn ) in C0 (X), then for each x ∈ X fn (x) is a Cauchy sequence of complex numbers, thus
it converges pointwise to some f (x). Furthermore on any compact subset of X the uniform
convergence fn (x) → f (x) implies continuity of f . Since X is locally compact and continuity
is a local property, it follows that f ∈ C(X), then using fn ∈ C0 it is not hard to see that
f ∈ C0 . (One way to check this is to use the net convergence characterization of continuity).
Now, given a function f ∈ C0 (X) we could approximate it by a convergent sequence of
compactly supported continuous functions on X using Urysohn’s lemma: let Kn = {|f | ≥
1/n} which is compact and sits inside Un = {|f | < 1/(n−1)} an open set. Then by Urysohn’s
lemma we may find φn a bump function that vanishes outside Un but equals 1 on Kn , it is
clear that φn f → f in the uniform metric.
3. L2 -Sobolev spaces on R. If k ≥ 0 integer we could define
H k := {f ∈ L2 : f, . . . , f (k) ∈ L2 }
1
For normed linear spaces, “ completeness” is equivalent to “every absolutely summable sequence is
summable”.
2.1. LINEAR FUNCTIONALS AND DUAL SPACES 25
be the space of L2 functions whose first k derivatives exists almost everywhere and are L2
integrable. H k is complete and we may alternatively define H k as the completion of the
space of locally compacted C ∞ functions on R under the norm
kf kH k = kf k2 + · · · + kf (k) k2
Using the Fourier transform, we could generalize this to allow for k fractional and even
negative, and we could also use Lp instead of L2 . We’ll revisit this in the future if time
permits.
4. If Y is a closed subspace of X then the quotient space X/Y (the space of equivalent
classes where x1 ∼ x2 if x1 − x2 ∈ Y ) is a normed linear space with norm
k[x]k = inf kx − yk
y∈Y
An equivalent definition is k[x]k = inf z∈[x] kzk. Note that closedness of Y is essential here to
ensure that k[x]k = 0 iff x ∈ Y .
5. If Y is a subspace of X then the closure Y of Y (with respect to the norm topology)
is another subspace of X. If this closure is the same as X then we say that Y is dense in X.
Note that this closure is not necessarily complete and Y is not the same as the completion
of Y under the norm. For instance we could take X = Y incomplete NLS, then the closure
of Y under the norm topology of X is the same as X, still incomplete.
k`(x)kY ≤ CkxkX
Then ` has an unique extension to a bounded linear map from X to Y and satisfies the above
estimate for all x ∈ X.
Proof. It is not hard to see that if such ` exists it has to be unique. To define `, fix x ∈ X.
Since D is dense in X there exists a sequence (xn ) in D that converges to x. We then define
`x = lim `xn . Note that this limit exists because `xn is a Cauchy sequence in Y which is a
complete space. One could easily show that the value of `x does not depend on the choice
of the sequence xn . Linearity and boundedness could be easily checked.
While working with singular integral operators on Lp spaces we typically invoke the
above theorem implicitly: these operators are explicitly defined only for a nicer dense subset
of functions (say sufficiently smooth and with sufficient decay), and so the theorem says
that as long as we could bound the operators on these dense subspace we could extend the
operator to all of the corresponding Lp and get the sameR 1 bound.
Example: the Hilbert transform Hf (x) = p.v. R y f (x − y)dy is defined for smooth
compactly supported functions on R which is dense in Lp . It turns out that kHf kp ≤ Ckf kp ,
1 < p < ∞, thus H extends to a bounded maps on Lp .
kT xkY ≤ kT kkxkX
2.1. LINEAR FUNCTIONALS AND DUAL SPACES 27
is in turn another Banach space. (For us Y = R (or C) with the distance norm). It is not
hard to see that this is a normed linear space, the main thing is to show completeness. Given
any Cauchy sequence (Tn ) in B(X, Y ) it is not hard to see that supn kTn k < ∞. Now for
any x ∈ X the sequence Tn x is a Cauchy sequence in Y therefore it converges (thanks to
completness of Y ) in Y , and we let T∞ x to be this limit. It is clear that T∞ is linear and
bounded kT∞ k ≤ supn kTn k < ∞. It remains to show that limn→∞ kTn − T k = 0.
Examples:
1. The dual space of a Hilbert space is itself.
2. If 1 < p < ∞ then the dual space of Lp (X, µ) is Lq (X, µ) where q is the conjugate
exponent 1/p + 1/q = 1. In particular,
Z
kf kLp (X,µ) = sup | f gdµ|
g:kgkq =1
If µ is a σ-finite measure(namely one could break the space into countably many subsets
where µ is finite) then the dual space of L1 (X, µ) is L∞ (X, µ). This may not be true without
the σ-finite assumption. ON the other hand, the dual space of L∞ (X, µ) generally speaking
not L1 (X, µ) (but there are examples when they are, say when X is a finite set with the
counting measure).
3. If X is a locally compact Hausdorff space, then the dual of C0 (X) is the space of
regular Borel measures with finite total mass. This result is one of Riesz’s representation
theorems and we will prove them later in the course.
2.1.3 Reflexivity
If X is a normed linear space we let X ∗ denote its dual and X ∗∗ denote the dual of X ∗ .
Definition 3 (Reflexive spaces). We say that X is reflexive if X = X ∗∗ (up to isomor-
phism).
In particular, a reflexive space has to be a Banach space to begin with, but certainly not
all Banach spaces are reflexive.
The interest in reflexive spaces is natural, since we always can isometricaly embed X into
X . To see this, fix x ∈ X. Let k.k∗ and k.k∗∗ be the norms on X ∗ and X ∗∗ respectively.
∗∗
b(`) := `(x) , ` ∈ X ∗ .
x
Examples:
1. Hilbert spaces are reflexive.
2. Lp are reflexive if 1 < p < ∞. As discussed above, generally speaking both L1 and
∞
L are not reflexive. For examples they are not reflexive when the underlying space is Rn
with Lebesgue measure or Z with counting measure.
3. A closed linear subspace of a reflexive space is also reflexive.
4. C[−1, 1] is not reflexive. Note that this space is separable but its dual is not (since
the dual of this space contains in particular all Borel measures). As we’ll see later, for a
reflexive space its dual space has to be separable too. One could prove this directly w/o
using separability.
√
so we could take C1 = m maxj kxj k1 . To show existence of C2 , note that the identity map
` : (X, k.k2 ) → (X, k.k1 ) (i.e. `x = x) is continuous since it is Lipschitz. The unit ball
2.2. (NON)COMPACTNESS OF THE UNIT BALL 29
in (X, k.k2 ) is compact, thus its image under this continous function is also compact, thus
bounded in k.k1 , therefore for some C2 > 0 we have
sup kxk1 ≤ C2
the sup is over x with kxk2 = 1, which is equivalent to the desired estimate.
In the infinite dimensional case, this is a theorem of Riesz.
Theorem 8 (Riesz). If X is an infinite dimensional NLS then the closed unit ball is not
compact with respect to the norm topology.
Proof. Our plan is to find a sequence of elements of unit vectors in X, x1 , x2 , . . . , such that
the distance kxi − xj k between any two elments of the sequence is uniformly larger than
1
0, for instance kxi − xj k ≥ 10 for all i 6= j. If that could be constructed it is clear that
no subsequence of (xn ) is Cauchy, thus no subsequence of this sequence is convergent and
therefore the closed unit ball is not (sequentially) compact.
To construct this sequence it suffices to show that if Y is a closed proper subspace of X
1
then one could find x ∈ X \ Y with kxk = 1 such that dist(x, Y ) ≥ 10 . Once this is proved
we could start with any point x1 of unit length and let Y be spaned by x1 (which is closed)
1
and then select x2 of unit length as above (in particular kx2 − x1 k ≥ 10 ), then reset Y to be
spaned by x1 , x2 (a closed subspace because it has finite dimension) and select x3 , etc.
1
Thus we only to show existence of x ∈ X of unit norm satisfying dist(x, Y ) ≥ 10 whenever
Y is a closed proper subspace of X. Let z ∈ X \ Y and let
d := dist(z, Y ) > 0 .
(if d = 0 then the closedness of Y would imply that z ∈ Y , contradiction.) Now for some
y ∈ Y we have kz − yk < 10d. Let x = z − y, it follows that dist(x, Y ) = dist(z, Y ) = d, thus
Exercises:
1. Prove that for a normed linear space, “completeness” of the norm is equivalent to
“every absolutely summable sequence is summable”.
2. Prove that for every x ∈ X a normed linear space it holds that kxk = sup`∈X ∗ : k`k=1 |`(x)|.
3. Prove that if K is compact Hausdorff then C(K) is complete with respect to the
sup norm. Use this to complete the proof that C0 (X) is complete if X is locally compact
Hausdorff without appealing to the one-point compactification trick.
4. Let S be a subset of X a normed linear space, and let Y be the linear span of S
(consisting of all finite linear combination of elements of S). Let L ⊂ X ∗ to be the set of all
` ∈ X ∗ that vanishes on S. Prove that z ∈ Y the closure of Y if any only if `(z) = 0 for
every ` ∈ L. (Hint: one direction should be easier, for the other direction you should use
the BLT theorem somewhere.)
5. Prove that all closed linear subspaces of a reflexive (Banach) space are reflexive. [Hint.
Use problem 4 at some point.]
Chapter 3
{x ∈ X : |b
y (x)| < }
Note that the maps y 7→ yb isometrically embedded Y inside X ∗ , the weak* topology is
smaller than or equal to the weak topology.
Certainly if X is reflexive then the weak* and weak topologies are the same.
Proof. Let X = Y ∗ . For simplicity assume that the spaces are over R. For each y ∈ Y let
Iy = [−kykY , kykY ] ⊂ R and consider
Y
I= Iy
y∈Y
31
32 CHAPTER 3. GEOMETRY AND TOPOLOGY ON NLS
x 7→ T x := (x(y))y∈Y ∈ I
(the fact that T x ∈ I follows because for x ∈ B we have |x(y)| ≤ kxkX kykY = kykY ). Note
that elements of T B are special because the coordinates of T x are actually related linearly.
Now, T is clearly linear and one to one and it embeds B into I, in fact it is not hard
to see that if we use the weak* topology on B and the inherited product topology on T B
then T is indeed a topological isomorphism between B and T B. We now show that T B is a
closed subset of I, which together with compactness of I will then implies that T B (hence
B) is compact.
Let z = (zy )y∈Y ∈ T B where closure taken inside I under the product topology. Then
we want to show that z is “linear and bounded by 1”, namely
Once that’s done it follows that the maps y 7→ zy defines a linear map on Y with norm at
most 1 and so z ∈ T B as desired.
To verify the above property, simply observe that there is a net in T B that converge to z
(unique because of Hausdorffness), and linearity relationship between coordinates is actually
preserved in the limit, so this property survives and z has it.
Examples:
1. If X is the space of regular Borel measures on R with the norm = total mass, then its
is the dual space of C0 (X) which is a separable space. Therefore the closed unit ball in X is
sequentially compact. In particular, given any sequence (µn ) of probability measures on R
there is a subsequence (µnk ) that converges vaguely to some Borel measures µ, i.e. for every
continuous function f that vanishes at ∞ it holds that
Z Z
lim f (x)dµn (x) = f (x)dµ(x)
n→∞
Theorem 11. If X = Y ∗ where Y is separable normed linear space, then the closed unit ball
B in X is sequentially compact in the weak* topology. In other words given any sequence
(xn ) in B there exists a subsequence xnk and x ∈ B such that
for every y ∈ Y .
Namely, since Y is separable the weak* topology is metrizable (i.e. it arises from some
metric on X), thus compactness and sequentially compactness are the same on this topology.
We could also prove this directly: let (yk ) be a dense subset of Y , then from the sequence
xn we could select a subsequence xnk such that y1 (xnk ) converges. Keep doing this and
use a diagonal argument we get a sequence xnk such that for any fixed j it holds that
limk→∞ yj (xnk ) exists. Using the fact that (yj ) is dense in Y it follows that for every y ∈ Y
the limit limk→∞ y(xnk ) exists, let this limit be z(y), it is clear that z is a linear functional
on Y and also bounded with kzk ≤ 1. This implies xnk converges weakly to z and element
of B.
Then Y is a closed linear subspace of X so is also reflexive, on the other hand Y is separable,
thus one could see that the closed unit ball inside Y is sequentially compact with respect to
the weak topology on Y . Thus there is a subsequence xnk and x∞ ∈ Y such that for every
bounded lienar functionals ` on Y we have
`(xnk ) → `(x)
Since every bounded linear functionals on X is also a bounded linear functional on Y , this
implies xnk converges weakly to x.
Definition 4 (Uniformly convex). We say that a normed linear space X is uniformly convex
if for each 0 < ≤ 2 there exists δ = δ() > 0 such that the following holds: If kxk = kyk = 1
and kx − yk ≥ then
x+y
k k≤1−δ
2
Note that this is a property of the norm: there may be equivalent norms on the same
space that is not uniformly convex.
Uniform convexity does not imply completeness. For instance we could take the rational
2 2
numbers with the usual distance. Then | x+y2
|2 + | x−y
2
|2 = |x| +|y|
2
which implies uniform
convexity.
Note that there are reflexive Banach spaces where it is not even possible to replace the
given norm with an equivalent norm that is uniformly convex, this is due P
to M. Day (BAMS,
1941). The idea is to use a vector valued Banach space with norm k.k = ( n kxj kp )1/p where
xn belongs to a reflexive Banach space Bn (which makes it reflexive); now for each j one
could still replace the norm on Bj by an equivalent norm that is uniformly convex, but as
n → ∞ these replacement couldn’t be done uniformly because the underlying equivalence
constants blow up if the Bj are carefully chosen.
Properties:
1. Hilbert spaces are uniformly convex, since in Hilbert spaces we have the paralellogram
law kx − yk2 + kx + yk2 = 2kxk2 + 2kyk2 .
2. C[−1, 1], L1 (R), L∞ (R) are not uniformly convex because they are not reflexive.
3. Clarkson(1936): if 1 < p < ∞ then Lp and `p spaces are uniformly convex.
For p ≥ 2 this holds essentially some form of parallelogram law holds.
0 0 0
(kx + ykp + kx − ykp )1/p ≤ 21/p (kxkp + kykp )1/p
3.3. ISOMETRIES BETWEEN NORMED LINEAR SPACES 35
0 0 0
where 1/p + 1/p0 = 1. For p ∈ (1, 2) we have a variant (kx + ykp + kx − ykp )1/p ≤
0
21/p (kxkp + kykp )1/p . 1
4. (Radon–Riesz, aka property (H)) If xn converges weakly to x in a uniformly convex
Banach space X, i.e. `(xn ) → `(x) for all ` ∈ X ∗ , and kxn k → kxk, then kxn − xk → 0. 2
5. A space is called uniformly smooth if its dual is uniformly convex.
6. A normed linear space is uniformly convex if the following holds for every bounded
sequences (xn ) and (yn ): if
kxn k2 + kyn k2 xn + y n 2
lim −k k =0
n→∞ 2 2
then limn→∞ kxn − yn k = 0. Below are two notions weaker than uniformly convex:
Local uniform convexity: A norm is locally uniformly convex if the following holds
for every x ∈ X and (xn ) in X: if
kxk2 + kxn k2 x + xn 2
lim −k k =0
n→∞ 2 2
then limn→∞ kx − xn k = 0. It is clear that this is a consequence of uniform convexity.
Strict convexity: A norm is strictly convex if the following holds for every x, y: if
kxk2 +kyk2
2
= k x+y
2
k2 then x = y. It is clear that this is a consequence of local uniform
convexity.
It can be shown that if X is a separable normed linear space over R then there exists an
equivalent locally uniformly convex norm.
kT x1 − T x2 kY = kx1 − x2 kX
for every x1 , x2 ∈ X. Note that we do not assume that T is affine (i.e. T (αx + βy) =
αT x + βT y, which is the same as linear if we impose T 0 = 0).
The Banach–Mazur distance: The distance is useful to compare two norms on Rn -
it is known that all norms are equivalent so the point is to be more quantiative.
1
Alternatively, use Hanner’s inequality which says that if k.k is the Lp or `p norm then
Analytic: Let X and Y be normed linear spaces. The multiplicative BZ distance is defined
to be
d(X, Y ) = inf{kT kkT −1 k}
the infimum is over all isomorphisms T : X → Y . Sometimes people also use log d(X, Y )
which satisfies the usual (additive) triangle inequality.
Geometric: given two bounded convex sets with nonempty interiors (aka convex bodies)
K and L in Rn that are symmetric (i.e. K = −K and L = −L) the Banach–Mazur distance
between them is
For Rn these two notions are equivalent. Basically given any norm on R the unit ball
is a symmetric convex body and conversely given any symmetric convex body K we could
construct a norm such that its unit ball is K, namely kxk = inf{t > 0 : x ∈ tK}.
Theorem 14 (F. John). Let `n2 denote Rn with the Euclidean norm. If X is a n-dimensional
NLS over R then
√
d(X, `n2 ) ≤ n
(This is actually sharp; equality holds for instance if X = `n∞ , for the geometric case take
the unit ball of this space.)
It follows from the theorem and the (multiplicative) triangle inequality that given any
two n-dimensional normed linear spaces over R it holds that
d(X, Y ) ≤ n
This result is also sharp upto a constant (Gluskin, 1981, a randomized construction).
Existence of nonlinear isometry: A natural question is whether non-affine isometry
exists (or equivalently nonlinear if we assume the normalization T 0 = 0)?
This is not true for complex NLS, even complex Banach spaces, see for instance the
conjugation mapping C to itself z 7→ z. We would need surjectivity, since otherwise the map
T : R → R2 mapping x to (x, sin(x)) is not linear but is an isometry if we equip R2 with the
L∞ norm k(x, y)k = max(|x|, |y|) (and distance norm in R).
It turns out that if the underlying normed linear spaces are over R then the answer is no.
Theorem 15 (Mazur-Ulam, 1932). If X and Y are NLS over R and T : X → Y surjective
and isometric with T 0 = 0 then T is linear.
If T 0 = 0 is not given the conclusion could be equivalently changed to “T is affine”,
namely T (αx + (1 − α)y) = αT (x) + (1 − α)T (y).
Main ideas of proof: Isometries are continuous so it suffices to show T (kx + my) =
kT (x) + mT (y) for all rational m, k, in fact suffices to show for k = m = 1/2. We construct
a nested sequence of subsets A1 ⊃ A2 . . . each of them is symmetric around x+y 2
such that
∩An = { x+y2
}. This construction will be invariant under surjective isometries, in particular
3.3. ISOMETRIES BETWEEN NORMED LINEAR SPACES 37
T x+T y
to get we could use the nested sequence T A1 ⊃ T A2 . . . , which are symmetric around
2
T x+T y
2
When that’s done we could take intersections to get T x+T
. 2
y
= T ( x+y
2
) as desired.
Now, let A1 contains all elements z of X such that kx − zk = ky − zk = kx−yk 2
. To
construct An+1 from An , simply let An+1 contains all z ∈ X such that for all w ∈ An it holds
that kw − zk ≤ 12 diam(An ), one could check that An+1 is still symmetric around x+y 2
and
diam(A1 )
diam(An+1 ) ≤ diam(An )/2 ≤ · · · ≤ 2n
→ 0 as n → ∞. Note that we need T X = Y
since we want T An+1 contains all z ∈ Y = T X such that “so and so” holds.
Exercises:
1. For any 1 < p < ∞ and any measure space (X, µ) prove that Lp (X, µ) is uniformly
convex (use the hints from the lecture notes). √
2. Show that the Banach Mazur distance satisfies d(`n∞ , `n2 ) ≥ n. [Hint: use the geomet-
ric formulation with corresponding convex bodies, and invoke a generalized√parallelogram
law.] (Note that together with F. John’s theorem this implies d(`n∞ , `n2 ) = n.) Then via
1 1
Holder’s inequality and the analytic formulation prove that d(`np , `nq ) = n| p − q | if 1 ≤ p, q ≤ 2
or 2 ≤ p, q ≤ ∞.
3. Prove F. John’s theorem for n = 2. [Hint. Use the geometric formulation, you may
want to use the fact that the Banach-Mazur distance is invariant under isomorphism of the
plane.]
4. Let X be a normed linear space. Let B = {x ∈ X : kxk ≤ 1} and B ∗∗ = {z ∈ X ∗∗ :
kzk∗∗ ≤ 1}. We know that there is a canonical map x 7→ x b that embeds B isometrically
linearly into B ∗. Prove that the image of B under this map is dense in B ∗∗ in the weak*
∗
topology of X ∗∗ (i.e. the minimal topology on X ∗∗ that makes all bounded linear functionals
constructed from elements of X ∗ continuous on X ∗∗ ). Use this fact to show that if B is
weakly compact then B = B ∗∗ and therefore X is reflexive.
5. Find two examples demonstrating that ”compactness” and ”sequentially compactness”
do not imply each other.
6. Let K be a convex subset of X a Banach space. Prove that if K is closed in the norm
topology then it is closed in the weak topology.
38 CHAPTER 3. GEOMETRY AND TOPOLOGY ON NLS
Chapter 4
A Hilbert spacep is a complete normed linear space where the norm arises from an inner
product kuk = hu, ui, and the inner product hu, vi is a bilinear form that satisfies the
following properties (we state the properties for Hilbert spaces over C):
(i) (u, v) ≥ 0, equality iff u = 0.
(ii) hu, vi = hv, ui.
(iii) hu, vi is linear in u (hence conjugate linear
R in v).
2
Examples: H = L (X, dµ) with hf, gi = X f (x)g(x)dµ; by Holder this is finite
Z Z Z
| f (x)g(x)dµ| ≤ ( |f | dµ) ( |g|2 dµ)1/2
2 1/2
X X X
39
40 CHAPTER 4. BASIS ON HILBERT SPACES AND BANACH SPACES
Theorem 16 (Riesz). For any ` ∈ H ∗ there is h` ∈ H such that `(x) = hx, h` i for all x ∈ H,
and the map ` 7→ h` is a surjective isometry (conjugate linear). Consequently H ∗ ∼ H.
Proof. Clearly ker(`) is a closed linear subspace of H, let Y = {y ∈ H : y ⊥ ker(`)}. Then
any x ∈ H has a unique decomposition into y + z with y ∈ Y and z ∈ ker(`). Sincd ` is
nonzero on Y so there is y0 ∈ Y such that `(y0 ) = 1. We now decompose each x ∈ H using
x = `(x)y0 + (x − y0 `(x)), now the first element is clearly in Y and the second is in ker(`)
since `(y0 ) = 1. It follows that H = span(y0 , ker(`)) and we could let h` = kyy00k2 , then
countable. Using this fact and Zorn’s lemma we could show that every Hilbert space has an
orthonormal basis.
3. If H is separable then we could define a natural isometry i : H → `2 (Z).
Schauder basis: A sequence E = (xn ) is a (Schauder) basis for a normed Pnlinear space X
if for every x ∈ H there is a unique scalar sequence αn such that limn→∞ kx− k=0 αk xk k = 0.
Equivalently, (xn ) is a Schauder Pbasis if two conditions hold: the closure of the linear
span of {xn } is the whole space and n=1 an xn = 0 iff an = 0.
Note that this is different from an algebraic basis (aka Hamel basis) which consists of
linearly independent vectors and any x ∈ X could be written as a finite linear combination of
these elements. Note that this could be defined on any linear spaces, topology (convergence)
is not needed, as we dont’ permit infinite sums. Also the order of the elements in a Hamel
basis is not important.
A Schauder basis is said to be an unconditional basis if it remains a basis even after
reordering. There are two other equivalent characterizations
Theorem 19. Let (xn ) be a (Schauder) basis in a Banach space (over R or C). The following
are equivalent:
(i) (xn ) is an unconditional basis. P
(ii) For any scalar sequence (αn ), if αn xn converges then it converges unconditionally
(i.e. it remains convergent even if we change the summation order).
(iii) There exists C > 0 finite such that
n
X n
X
k k αk xk k ≤ Ck α k xk k
k=1 k=1
uniformly over n and all sequences n with |n | ≤ 1 and all scalar coefficients α1 , . . . , αn .
Property:
1. For NLS, existence of Schauder basis means the space is separable.
2. Every orthogonal basis in a separable Hilbert space is an unconditional Schauder basis.
Examples: √
1. Fourier basis: L2 [0, 1] has (e2πinx / 2π)n∈Z as an unconditional Schauder basis. This
remains a basis for Lp [0, 1] 1 < p < ∞ but not unconditional for p 6= 2.
2. Haar basis: 1[0,1] and hI (x) = √1 (1Il − 1Ir ) indexed by dyadic subintervals I ⊂ [0, 1]
|I|
2 p
is a basis for L and all L [0, 1] with 1 ≤ p < ∞, note that p could be 1 here. This is an
unconditional basis if 1 < p < ∞.
3. Orthogonal polynomials: If w(x), x ∈ R has sufficiently fast decay (say subexpo-
nential - Bernstein’s theorem) then the orthogonal polynomials with respect to dµ = wdx is
42 CHAPTER 4. BASIS ON HILBERT SPACES AND BANACH SPACES
a basis for L2 (R, µ). Here pn (x) = an,n xn + · · · + a0,n where an,n > 0 and
Z
pn (x)pm (x)dµ = δmn
Proof of Theorem 19. It is clear that (i) and (ii) are the same. So we only show equivalence
to (iii).
We first observe that the following are equivalent for any given sequence (xn ) in a Banach
space X: P
(a) the series n xn converges unconditionally; P
(b) for any > 0 there exists n = n() such that k k∈M xk k < for any finite subset
M ⊂ (n, ∞) of the integers.
Note that the direction (b) → (a) is a consequence of the completeness
Pn of the space.
(Basically (b) implies that for any permutation of N the sequence k=1 xσ(k) would be a
Cauchy sequence).
For the other direction (a) → (b), assume towards a contradiction that there is an > 0
such that no such n could be found,
P thus we could find a sequence of sets M1 , M2 , ... such
that maxMj < minMj+1 and k k∈Mj xk k > . It is not hard to build a permutation σ of N
such that each Mj appears as a consecutive block of σ(1), σ(2), . . . , thus nk=1 xσ(k) is not a
P
Cauchy sequence in X so won’t be convergent, contradiction. P
Using
P this observation we claim
P that if a given series n xn is unconditionally convergent
then n xσ(n) is the same as n xn for all permutation of N. Using this fact it is not hard
to see that (i) and (ii) are equivalent. To see this claim,
P just note that forPany σ there is
some M such that σ(M + 1), · · · > n(), therefore k k>M xσ(k) k ≤ and k k>n() xk k ≤ ,
and by canceling out the common terms it follows that
X X X
k xk − xσ(k) k = k xσ(k) k <
k≤n() k≤M k≤M :σ(k)>n()
P P
thus k k xσ(k) − k xk k < 3 for all > 0, implying the desired claim.
Now we show that (ii) and (iii) are equivalent. In fact for simplicity we will only show
the equivalence to the version with = ±1. P
It is clear that (iii) implies (ii): suppose that Pk αk xk converges and (iii) holds true.
Then for any > 0 there exists n = n() such that k k>n αk xk k < /(2C). Given any finite
A ⊂ (n, ∞) by choosing the sign sequence that equals 1 P on A, equals 0 on P {1, 2, . . . , n} and
equals ±1 constantly on N\A and using (iii) P we obtain k k∈A αk xk k ≤ 2C k>n αk xk k < .
Thus by the observation above the series k αk xk converges unconditionally.
Now, to show that (ii) implies the sign sequence version of (iii), we will use the following
lemma
P P
Lemma 3. Let Pn be the projection Pn x = k≤n αk xk if x = k αk xk is the expansion into
the Schauder basis (xn ) for x. Then supn kPn k < ∞.
4.3. SCHAUDER BASIS 43
We first prove the above implication using the lemma. The proof of the lemma will be
discussed in the next class, after we have introduced several basic tools in Banach spaces,
such as the principle of uniform boundedness and the open mapping theorem. Now, it suffices
to show that there exists N > 0 and a scalarPCN > 0 such that forP every sequence of scalar
αj and signs j and n ≥ N it holds that k j=N j αj xj k < CN k nj=1 αj xj k. Indeed, the
n
desired claim would follows from the fact that given any N and any sign sequence 1 , . . . it
holds uniformly over n ≥ N that
N
X n
X
k j αj xj k .N k α j xj k
j=1 j=1
To see this, simply use Lemma 3 to obtain kαj xj k ≤ k(Pj−1 −Pj )( nk=1 αk xk )k ≤ Ck nk=1 αk ck k
P P
and then use the triangle inequality.
Now assume towards a contradiction that for every N > 0 one can not find such CN .
We then construct a sequence n1 < n2 < . . . and Aj ⊂ [nj , nj+1 − 1] andPβ1 , β2 , . . . re-
cursively as follows, with the following property: for any j it holds that k k∈Aj βk ak k ≥
j 2 k nj ≤k≤nj+1 −1 βk ak k.
P
Suppose that we have chosen n1 < · · · < nj and A1 , . . . , Aj−1 and β1 , . . . , βnj −1 satis-
fying the above requirements. Then given C > 0 by the assumption there exists nj+1 and
β1 , . . . , βnj+1 −1 and B ⊂ [nj , nj+1 ) such that
X X X
k β k xk − βk xk k ≥ Ck β k xk k
k∈B k∈[nj ,nj+1 )−B 1≤k≤nj+1 −1
from here it is clear that either Aj = B or Aj = [nj , nj+1 ) − B will work, with αk = βk for
k ∈ [nj , nj+1 ). This completes the selection of Aj and nj and αj .
P Pnj+1 −1
Now by rescaling αk ’s we may assume that k k∈Aj αk xk k = 1 while k k=n j
αk xk k ≤
2
1/j . Now, given any m1 < m2 we let ns < · · · < nt be the elements of (nj ) inside [m1 , m2 ),
then using the uniform boundedness of Pn we obtain
m2 t
X X X 1
k α k xk k . k αk x k k .
k=m1 j=s−1 nj ≤k<nj+1
s
44 CHAPTER 4. BASIS ON HILBERT SPACES AND BANACH SPACES
P
and clearly if m1 → ∞ then so is s, thus k αk xk converges. On the other hand it does
not converge unconditionally since it violates the second property of the initial
P observation:
given any n we could select j such that nj > n, and so Aj ⊂ (n, ∞) and k k∈Aj αk xk k = 1.
Exercise:
1. Prove that (e2πinx )n∈Z is an unconditional Schauder basis for L2 [0, 1].
2. Prove that the Haar functions are orthogonal in L2 [0, 1] and form a Schauder basis
for L1 [0, 1] but not unconditional in L1 .
3. (Radon–Nikodym) Let µ and ν two nonnegative σ-finite measures on the same σ
algebra (X, A) such that the following holds: for every measurable A if µ(A) = 0 then
ν(A) = 0. Show R that there exists g nonnegative such that for every measurable E it holds
that ν(E) = E gdµ.
Chapter 5
Here we will prove several basic results about bounded linear maps between Banach spaces:
the principle of uniform boundedness, the open mapping theorem, and the closed graph
theorem, and applications. We then focus on Lp spaces.
45
46 CHAPTER 5. BOUNDED MAPS ON BANACH SPACES
Proof. We first show that for a linear operator T : X → Y (only need normed linear spaces
X, Y ) T is bounded if any only if T −1 (kykY ≤ 1) has nonempty interior in X.
Certainly if T is bounded then the set T −1 (kykY ≤ 1) contains {kxkX ≤ 1/kT k} so
nonempty interior.
Conversely, if T −1 (unit ball in Y ) contains an interior point x0 , then for some > 0 it
holds for every kxkX ≤ that kT (x0 + x)k ≤ 1. It follows from the triangle inequality that
kT xk ≤ 1 + kT x0 k for such x, thus kT k ≤ 1+kT x0 k .
Therefore in order to show P.U.B it suffices to show that the set
\
B1 = Tα−1 (kyk ≤ 1)
α∈A
has an interior point. In fact it suffices to show thatSfor some n > 0 the (closed) set Bn
(replace the radius by n) has nonempty interior. But Bn = X because supα kTα xk < ∞.
Thus the desired claim follows from the Baire category theorem.
it is not hard to see that this is a norm and kxk ≤ kxk1 . We will show that Pn is bounded
in (X, k.k1 ) and the norm k.k1 is actually equivalent to k.k, these two facts will take care of
the lemma. Now the boundedness of Pn with respect to the new norm is clear
Now we show that (X, k.k1 ) is complete. Once we did that, the identity map from
(X, k.k1 ) → (X, k.k) is a bijective bounded linear map thus by the open mapping theorem
it is boundedly invertible, thus the two norms are equivalent and completes the proof of the
lemma. Now, let X 0 be the completion of X under k.k1 , then it is not hard to see that P xn is
0 0
still a Schauder basis for (X , k.k1 ) (see the homework). So if a ∈ X we have P a = j α j xj
convergence
P in k.k1 , but this series also converges in k.k too because k m<j≤n αj xj k ≤
k m≤n αj xj k1 → 0 and X is complete. So, let b ∈ X be its limit under k.k. Now, it is clear
that b = a because
X X
kb − αj xj k1 = sup kPm (b − αj xj )k = sup k(Pm − Pn )bk → 0
m m≥n
j≤n j≤n
if n → ∞.
5.6 Lp spaces
In this section we conduct a case study of Lp spaces, which are the most fundamental type
of Banach spaces.
5.6.1 Separability
Recall that if (X, A, µ) is separable then Lp (X, A, µ) is separable for 1 ≤ p < ∞; this is part
of the homework.
5.6.2 Duality
We’ll use Hilbert space techniques to show that
Theorem 24. For 1 < p < ∞ the dual of Lp (X, µ) is Lq (X, µ) where 1/p + 1/q = 1.
If furthermore the measure µ is σ-finite then the dual of L1 is L∞ .
Proof. We first consider the simpler case when µ is σ-finite, in this case we will show the
result for all 1 ≤ p < ∞.
If µ is not necessary σ-finite we will need p > 1. Let ` ∈ (Lp )∗ . Suppose that E ⊂ X
measurable such that µ|E is σ-finite. Clearly ` is also a bounded linear functional on Lp (E, µ)
whose norm does not exceed k`k. Thus, by the σ-finite case, for every f ∈ Lp (E) there is
gE ∈ Lq (E, µ) such that Z
`(f ) = f gE dµ
E
kgE kq ≤ k`k
Now if E ⊂ E 0 both measurable and σ-finite wrt µ then it is clear that gE = gE 0 on E.
5.6. LP SPACES 49
Now, using the fact that a countable union of σ-finite sets is σ-finite one could find F
σ-finite wrt µ such that kgF kq = supE σ−f inite kgE kq. We will show that
Z
`(f ) = f gF dµ
Rfor every f ∈ Lp (X). Observe that g|A\F = 0 for every A σ-finite, because otherwise
q
|gA\F | dµ > 0 and therefore kgA∪F kq > kgkF contradiction (note that here is where we
need p > 1 so that q < ∞). In particular let A = {|f | > 0} which is σ-finite wrt µ and so
Z Z
`(f ) = f gA dµ = f gF dµ
We now focus on the σ-finite case. For simplicity we will assume µ(X) < ∞, otherwise
we could decompose X into countably many subsets where µ is finte and apply this argument
to each subspace and add things up: because of continuity of `, for each f ∈ Lp (X) we have
XZ
`(f ) = lim f gk dµ
n→∞ Xk
k≤n
ν(E) = `(1E )
thus Z
( |g|q 1|g|≤K dµ)1/p ≤ k`k
mapping from measureable functions on (X, µ) to measurable functions on (Y, ν). At the
beginning the operator T may not be defined for all f , as the integral may not be convergent.
The adjoint of T is Z
T ∗ g(x) = K(x, y)g(y)dσ(y)
X
Think of this as a generalization of the matrix in finite dimensional setting. (integration
is like adding over indices and K gives the matrix entries.) (This turns out to be the case
if say X = {1, . . . , n} and Y = {1, . . . , m} with the counting measures.) The function K is
called the kernel (assumed measurable etc.).
Remarks: The dual estimates hold for T ∗, namely we also have the same L2 → L2
esimates and
Z Z
1/p 0
kT kLp0 →Lp0 ≤ [sup |K(x, y)|dσ(y)] [sup |K(x, y)|dµ(x)]1/p
x X y X
These turn out to be the easiest cases. For the L1 case this follows from
Z Z
kT f kL1 (Y,ν) ≤ |K(x, y)||f (x)|dµ(x)dσ(y)
Y X
Z Z
= |f (x)|[
|K(x, y)|dσ(y)] dµ(x)
X Y
Z
≤ kf k1 sup[ |K(x, y)|dσ(y)]
x X
∞
Similarly for the L case we have, for each y ∈ Y
Z
|T f (y)| ≤ |K(x, y)||f (x)|dµ(x)
X
Z
≤ kf k∞ sup |K(x, y)|dµ(x)
y X
To get the case 1 < p < ∞ of the first estimate, we will need Riesz-Thorin interpolation
theorem (aka convexity theorem).
The proof of this is an application of the maximum principle: if H(0) = H(1) = 1 then
the claims follows by the maximum principle, in the general case modify h by a suitable
exponential factor to reduce to this setting.
52 CHAPTER 5. BOUNDED MAPS ON BANACH SPACES
We now set up the bounded analytic function h(z). For Re(z) ∈ [0, 1] let pz and qz be
defined by
1 1 1 1 1 1
( , ) = (1 − z)( , ) + z( , )
pz qz p0 q 0 p1 q1
0
Then pz and qz are analytic functions. For any f ∈ Lp and g ∈ Lq (1/q 0 + 1/q = 1) we define
Z
h(z) = T (fz )gz dσ
Y
p q
where fz (x) = |f (x)| pz −1 f (x) and gz (x) = |g(x)| qz −1Rg(x). It is not hard to see that h(z)
is bounded analytic for Re(z) ∈ [0, 1] and h(α) = T f gdσ, therefore using the duality
characterization for Lp we obtain
M= sup |H(α)|
f ∈Lp ,g∈Lq0
0
Note that everything depends on f, g where f ∈ Lp and g ∈ Lq . We normalize kf kp =
kgkq0 = 1, then it is not hard to see that
|H(0)| ≤ M0 , |H(1)| ≤ M1
therefore the desired claims follows from Hadarmard three lines lemma.
Exercises:
1. Let X be a Banach space. Prove that separability of X ∗ implies separability of X.
2. Prove that the measure space (X, A, µ) is separable if any only if L2 (X, A, µ) is
separable. Is this true if we replace 2 by any fixed p ∈ (1, ∞)? Use this to construct a finite
measure space (X, µ) (µ(X) < ∞) such that L2 (X, µ) is nonseparable.
3. Prove that if X is a compact metric space then C(X) with the sup norm is separable.
4. Let X be a normed linear space and assume that (xn ) is a Schauder basis, furthermore
the canonical projection Pn into the first n vectors is bounded for each n ≥ 1. Show that (xn )
remains a Schauder basis for the completion of X. [Hint: show that Pn extends boundedly
to the completed space and use Pn to find the linear expansion/proving required properties
for Schauder basis.]
5. Let ` be a bounded linear functional on Lp (X, µ) where µ(X) < ∞. Show directly
(without using the duality theorem) that ` could be written as a linear combination of finitely
many positive linear functionals on Lp (X, µ). Here 1 ≤ p < ∞.
6. Show that every NLS is isomorphic to a subspace of the space of bounded continuous
function on some complete metric space Y . 1
1
if the given space is separable then one could take Y to be [0, 1] - this is the Banach-Mazur theorem.
Chapter 6
Recall that the dual space of a normed linear space is a Banach space, and the dual space
of Lp is Lq where 1/p + 1/q = 1 if 1 < p < ∞. If the underlying measure space is σ-finite
then the dual of L1 is L∞ . What about the dual of L∞ ? Or its subspace Co (X) and Cc (X)
where say X is locally compact Hausdorff. In this chapter we discuss several representation
theorems related to these themes.
53
54 CHAPTER 6. BOUNDED CONTINUOUS FUNCTIONS AND DUAL SPACES
Theorem 27 (Riesz). Let X be a LCH. Then the dual of Co (X) is M (X) the space of
complex valued Radon measures (locally finite regular Borel measures) on X. Namely the
following map is an isometric isomorphism between this dual and M (X)
Z
µ 7→ `µ (f ) = f dµ
Pm
and k`µ k = kµk the total variation of µ defined by sup j=1 |µ(Aj )| supremum over all
collection A1 , . . . , Am of disjoint subsets of X.
Step 1: We first show that if ` is a bounded positive linear functionals on Co,R (X), the
space of real valued continuous functions on X that vanish at ∞, then there is a Radon
measure µ such that for any f ∈ Co,R (X) it holds that
Z
`(f ) = f dµ
To see this, we would like to define µ by µ(E) = `(1E ), however 1E is not continuous therefore
one can not apply ` to this function. To get around this we may define for each open set
U ⊂X
Defintion: µ(U ) is defined to be the supremum over `(f ) where f ∈ Co (R) and the
support of f is inside U and 0 ≤ f ≤ 1 pointwise.
(Note that we may not be able to do this if U wasn’t open since such a continuous
function may not exists; for open U the existence of f is due to Urysohn’s lemma).
Now we obtain a premeasure on open sets which is a ring of sets, thus by Caratheodory
we may exend µ to the sigma algebra generated by these sets, which are exactly the Borel
sets. R
We want to show that `(f ) = f dµ for every f ∈ Co (X). Without loss of generality
assume 0 ≤ f ≤ 1. Since f vanishes at infinity the sets Ak = {f ≥ k/n} are compact, and
we may find continuous functions fk such that
consequently Z
1 1 1
|`(f ) − f dµ| ≤ µ(A0 ) ≤ µ(supp(f )) = O( )
n n n
56 CHAPTER 6. BOUNDED CONTINUOUS FUNCTIONS AND DUAL SPACES
≤ sup{`(g) : 0 ≤ g ≤ (1 + )f }
≤ (1 + )`(f )
We now show regularity properties for µ. (It is clear that µ is finite on compact set using
Urysohn’s lemma.) Outer regularity of µ follows from construction using premeasure, and
it remains to show inner regularity. Let E be Borel, it suffices to show that
now using outer regularity it suffices to do this for E open. Let U be an open set, we have
≤ sup{µ((supp(f )) . . . }
≤ sup{µ(K) : K ⊂ E compact}
Step 2: We now reduce the desired result to the positive setting. This is done via writing
bounded linear functionals on Co (X) as a linear combination of positive linear functionals. To
see this note that without loss of generality it suffices to consider bounded linear functional
on C0,R (X) real valued members of C0 (X). Let `+ be defined by
and let `− = `+ − `, it is not hard to check that both `+ and `− are positive linear functionals
on C0,R (X).
Note that there are also versions for complex valued functions, in which case the same
conclusion holds if we assume further that P is closed under conjugation p ∈ P then p ∈ P .
There are also versions for C0 (X) and C0,R (X) where X is locally compact Hausdorff.
The original Weiertrass approximation theorem is for X being compact intervals, which
implies that polynomials are dense inside continuous functions. Applications in probablity
(moment problems etc.)
Proof: Since P is also an algebra that separate points, without loss of generality we may
assume that P is closed. We divide the proof into two steps
Step 1: We will show that P is a lattice, namely if f, g ∈ P then so are max(f, g) and
min(f, g).
Step 2: Using the lattice property of P , we will show that if f ∈ CR (X) be such that for
every x 6= y in X there exists h ∈ P such that h(x) = f (x) and h(y) = f (y), then f ∈ P .
Combining these two steps, we prove the theorem as follows. Given any x 6= y consider
the algebra (h(x), h(y)), h ∈ P . Clearly this is a subalgebra subspace of R2 and it can’t be
{(0, 0)} since P separate points. Thus it could be either 0 × R or R × 0 or R2 . Now in the
last case by step 2 any f ∈ CR (X) must be in P as desired. In the first case for instance,
h(x) = 0 for all h ∈ P and so we must have {h(y), h ∈ P } = R for any other y. So by Step
2 again it follows that any function f such that f (x) = 0 must be in P as desired.
Proof of step 1: since max(f, g) and min(f, g) could be written as linear combination of
|f + g| and |f − g| and f and g, therefore it suffices to show that if f ∈ P then so is |f |. The
idea is to approximate |f | with a polynomial of f uniformly. Indeed, clearly f is bounded
therefore without loss of generality assume |f | ≤ 1, it then suffices to show that |x| could be
approximated
√ uniformly by polynomials on [−1, 1]. To see this use the Taylor expansion of
1
1 − t = 1 − 2 t + . . . which converges uniformly on 0 ≤ t ≤ 1, then write
√ p 1
|x| = x2 = 1 − (1 − x2 ) = 1 − (1 − x2 ) + . . .
2
Proof of step 2: Let > 0. It suffices to show that there exists g ∈ P such that
|g(x) − f (x)| ≤ uniformly over x ∈ X. Assume that for every x ∈ X there exists gx ∈ P
such that |gx (x) − f (x)| < and gx (y) ≤ f (y) + for every y ∈ X. Then the desired g could
be constructed as follows. By continuity for each x ∈ X there exists Ux ⊂ X neighborhood
of x such that supUx |gx − f | < . By compactness of X one could refine the collection
Ux , x ∈ X to a finite subcollection Ux1 , . . . , Uxm , and we simply let
g = max(gx1 , . . . , gxm )
Recall that C(X) is not a normed linear space when X is not compact. On the other hand
we could use semi norms on C(X): given any compact K ⊂ X let k.kK be the sup nom on
K. This family of seminorms determine C(X): together they provides a lot of properties so
that various theorems (for NLS) remains valid here.
59
60 CHAPTER 7. LOCALLY CONVEX SPACES
neighborhood base at 0 could be taken to consists of all open sets of the form
If one assume that the (countable) family is separated then the above pseudo metric is
actually a metric. If this metric is furthermore complete then the given space is called a
Frechet space.
Examples: Recall that C(X) is a locally convex space with seminorms given by ρK (f ) =
supx∈K |f (x)| where K ⊂ X is compact. Other examples are
(i) the space of Schwartz functions on Rn : These are functions that are C ∞ and their
derivatives decay faster than any polynomial. One could use ρα,β (f ) = |x|α |Dβ f (x)| where
α and β are nonnegative integer multi-indices.
(ii) if `α , αA is a family of linear functionals on X then the minimal topology such that
`α are continuous and linear operations (addition/scalar multiplication) are continuous is
called the weak topology induced by this set of linear functionals. Examples of these (for
normed linear spaces) are the weak topology and the weak* topology. It can be shown that
if ` is a linear functional that is continuous with respect to this topology it must be a finite
linear combination of the given linear functionals.
7.2. THE HYPERPLANE SEPARATION THEOREM 61
(iii) Given any open set Ω ⊂ Rn the space D(Ω) consisting of compactly supported
infinitely smooth functions on U is also a locally convex space and actually complete (recall
that the version without infinite smoothness, i.e. Cc (U ), is not complete).
(Note that this generalizes the usual equivalence of continuity and boundedness of linear
maps between normed linear spaces.)
Theorem 29. There exists a continuous linear functional ` on X such that supx∈K `(x) ≤
`(y). If furthermore K is closed then this inequality could be taken strict.
since 0 is an interior point of K it is clear that ρK (x) < ∞ for any x ∈ X. Furthermore ρK
is convex and positive homogeneous, and since y 6∈ K we have ρK (y) ≥ 1 ≥ ρK (x) for every
x ∈ K. Let ` be defined on the one dimensional subspace spanned by y using `(y) = 1. Then
|`(z)| ≤ ρK (z) inside this subspace, so by Hahn Banach we may extend ` to all of X such
that |`(x)| ≤ ρK (x) for all x. In particular `(x) is continuous since ρK is continuous at 0.
Now, if K is closed then one could see that ρK (y) > 1 ≥ supx∈K ρK (x) and we could
therefore obtain a strict inequality.
For a locally convex Hausdorff space over R, it follows from the above theorem that
for any two distinct points x 6= y there is a continuous linear functional ` on X such that
`(x) 6= `(y). To see this, by Hausdorffness and local convexity we could find a convex open set
A such that x ∈ A while y ∈ X \ A. Since A is closed convex we could apply the hyperplane
separation theorem and get a continuous linear function ` such that supz∈A `(z) < `(y). In
particular `(x) < `(y).
62 CHAPTER 7. LOCALLY CONVEX SPACES
Again the set of maximum of ` on K is a proper subset of K and also an extreme subset, and
7.4. INDUCTIVE LIMIT AND WEAK SOLUTIONS 63
this set is also closed and disjoint from E. So by repeating the above argument one could
find one extremal point of K inside this set, which is therefore not inside E, a contradiction.
Examples: Let X be a compact Hausdorff space and consider CR (X) (we could do locally
compact Hausdorff too with CR,0 (X)) and let A consists of all positive linear functional on
C(X). Let A be the subset of A with `(1) = 1, it is not hard to see that A is convex and its
extreme pooints are the point evaluation linear functional ex f = f (x).
A weak solution to a PDE is a distributional solution in the above sense. If this distribution
arise from some sufficiently smooth functions then we say it is a classical solution.
64 CHAPTER 7. LOCALLY CONVEX SPACES
Part II
65
Chapter 8
Let B denote the space of all bounded linear operators on some given Banach space X
over C. The analysis here works for more general settings, say B could be a unital Banach
subalgebra of this space (i.e. an associative subalgebra that contains a unit and is a complete
subspace with respect to the operator norm).
I + (N − I) + (N − I)2 1 + . . .
Definition: The resolvent set of M in B consists of all complex number λ such that
λI − M is invertible in B. The resolvent set is denoted by ρ(M ) and the spectrum of M is
σ(M ) := C \ ρ(M ).
Properties:
1. ρ(M ) is open in C.
(This is a consequence of the above observation.)
2. The resolvent function (λI − M )−1 is an analytic function of λ on ρ(M ).
This is because the above observation also shows that the resolvent could be expanded as
a power series around each point λ in the resolvent set ρ(M )
∞
X
−1
(λ − h − M ) = (λ − M )n−1 hn
j=0
67
68 CHAPTER 8. ELEMENTARY SPECTRAL THEORY
Closedness is clear, boundedness follows from the fact that if λ is large enough in modulus
then λ−1 M has small norm and so (1 − λ−1 M ) is invertible and so λ ∈ ρ(M ). To show that
σ(M ) is nonempty assume towards a contradiction that it is. Then consider the contour
integral along CR = {|ξ| = R} Z
(λ − M )−1 dλ
CR
clearly analyticity of the resolvent operator implies that this would be 0. On the other hand
we know that (λ − M )−1 has the Laurent series expansion at ∞
∞
X
−1
(λ − M ) = M n ξ −(n+1)
n=0
therefore the contour integral will gives the residue term i.e. the term with n = 0 and we
get I, contradiction.
To show the identity for the spectral radius, by elementary complex analysis it follows
that |σ(M )| ≥ lim sup kM k k1/k . So it suffices to show that |σ(M )| ≤ |M k |1/k for all k ≥ 1.
To see this, observe that
X k−1
hX ih X i
n −(n+1) j −(j+1) k −k m
k M ξ k≤ kM k|ξ| (kM k|ξ| )
n≥0 j=0 m≥0
this follows from kM mk+j k ≤ kM j kkM k km . It follows that if |ξ| > kM k k then the se-
ries converges absolutely and therefore ξ ∈ ρ(M ); this completes the proof that |σ(M )| =
lim kM k k1/k .
note that by the Cauchy theorem this is independent of the choice of the contour. It turns
out that this is consistent with the usual polynomial case and furthermore
σ(f (M )) = f (σ(M ))
(also known as the spectral mapping property). We also have the resolvent identity
which is useful to show that the functional calculus maps the algebra of analytic functions
on open sets containing σ(M ) into B is a homomorphism.
Also, if g is analytic on an open set containing f (σ(M )) and h = g ◦ f then h(M ) =
g(f (M )).
Spectral Projections: Assume the spectrum σ(M ) could be decomposed into an union
of n disjoint closed components σ1 ∪ . . . σn .
Let Cj be a contour R in ρ(M ) −1
that winds once around each point of σj but not other
components, and Pj = Cj (ξ − M ) dξ.
Then Pj are disjoint projections Pj2 = Pj and Pj Pk = 0 if j 6= k, and
X
Pj = I
j
and if σj 6= ∅ then Pj 6= 0.
1. Shifts
Let X = `2 consiting of x = (a0 , a1 , . . . ) such that |aj |2 < ∞. The right and left shifts
P
R and L
Rx = (0, a0 , a1 , . . . , ) , Lx = (a1 , a2 , . . . )
Now LR = I but RL 6= I so these linear bounded maps are not invertible. The spectrum of
R and L consists of all points in the unit disk {λ ∈ C : |λ| ≤ 1}. Now R0 = L and L0 = R
as adjoint of each other.
The idea is the spectral radius of L is 1Rso spectrum contained inside the disk.
2. The Fourier transform F f (ξ) = f (x)e−2πixξ dξ. Note that√F 4 = I therefore the
spectum of F is a subset of {x ∈ C : x4 = 1} which consists of ±1, ± −1. Let Hn denotes
2
the orthgonal polynomials wrt to e−x dx
√
Z
2
Hm Hn e−x dx = π2n n!δmn
70 CHAPTER 8. ELEMENTARY SPECTRAL THEORY
2 /2 √ 2
then it could be shown that F [e−x Hn (x)] = (− −1)n e−ξ /2 Hn (ξ). This follows from
2 /2
X
exp(−x2 /2 + 2xt − t2 ) = e−x Hn (x)tn /n!
n≥0
Rx
3. Volterra integral operator Let X = C([0, 1]) and let T f (x) = 0 K(x, t)f (t)dt,
also known as the Volterra integral operator, here K is continuous. The spectrumn of T
consists of only one point λ = 0. This follows by computing T n f , for instance if K ≡ 1 then
1
R x
T n f (x) = (n−1)! 0
(x − t)n−1 f (t)dt and show that kT n f k ≤ kf k/n! in the sup norm, and use
the spectral radius identity.
4. Diagonal multiplication Let X = `p (Z+ ) and M acts diagonally (M x)n = λn xn ,
then the spectrum of M is the closure of {λn }.
Proof. : It suffices to show, using the open mapping theorem, that T ∗ : X ∗ → X ∗ bounded
and injective and onto.
Injectivity: If T ∗ ` = 0 for some ` ∈ X ∗ then we will show that ` = 0.
hT x, `i = hx, T ∗ `i = 0
for every x, so using the fact that T is onto we obtain `(X) = 0 so ` = 0.
8.4. ADJOINT OPERATORS AND SPECTRUM 71
and so
hz, `i = y, [T ∗ − −λ]−1 `
73
74 CHAPTER 9. COMPACT OPERATORS
To see property (iv), first note that T xn converges weakly to T x. To see this, let ` ∈ X ∗ ,
then `(T xn ) = (T ∗ `)xn converges to T ∗ `x which is the same as `(T x). It then follows that
any strongly convergent subsequence of T xn has to converge to T x. Now since xn converges
weakly to x in X it follows that xn is a family of bounded operator on X ∗ that is uniformly
bounded pointwise, consequently by the principle of uniform boundedness supn kxn k < ∞,
thus (T xn ) is precompact, so any subsequent contains a Cauchy subsequence; thus by a
routine argument it follows that T xn is convergent to T x.
(v) If Tn is a sequence of compact operators from X to Y and Tn − T k → 0 as n → ∞
for some T : X → Y bounded linear, then T is compact.
To see property (v), note that for any > 0 one could choose N large such that kTN −T k <
/2. Now, we could use finitely many /2 balls to cover TN B1 . The balls with the same
centers will then cover T B1 . (Recall that B1 is the unit ball in X.)
As a corollary, we have
Corollary 3. If T : X → Y is the limit of a sequence of finite rank operators in the norm
topology then T is compact.
The converse direction of this corollary is not true in general (for Banach spaces X and
Y ), the first construction of counter examples is due to P. Enflo (’73). However, if Y is a
Hilbert space then the converse is true.
Theorem 32. Any compact operator T : X → Y where X is Banach and Y is Hilbert can
be approximated by a sequence of finite rank operators.
Proof.
SMn We sketch the main ideas of the proof. For every n the set T B1 is covered by
1
B
k=1 Y (y ,
k n ). We may assume that Mn is an increasing sequence. Let Pn be the projection
onto the span of y1 , . . . , yMn , which is clearly a finite rank operator, thus Pn T is also finite
rank. It remains to show that kT − Pn T k = O( n1 ). We observe that kPn y − yk k ≤ ky − yk k
for every 1 ≤ k ≤ Mn and every y ∈ Y , since projection are contractions. It follows that
and clearly given any y ∈ B1 there is a k such that kT y−yk k ≤ 1/n, therefore kPn T y−T yk ≤
2
n
for every y ∈ B1 , thus kPn T − T k ≤ 2/n as desired.
Theorem 33 (Schauder). Let T : X → Y be a bounded linear operator between Banach
spaces X and Y . Then T is compact if any only if T ∗ : Y ∗ → X ∗ is compact.
Proof. It suffices to show the forward direction, namely if T is compact then T ∗ is also
compact. For the other direction, apply the forward direction it follows that T ∗∗ is compact
from Y ∗∗ to X ∗∗ , and by restricting T ∗∗ to X we obtain T therefore T is also compact.
Now, assume that T is compact. Given any sequence `n ∈ X ∗ with k`n k ≤ 1 we will
show that (T ∗ `n ) has a Cauchy subsequence T ∗ `nj , in other words given any > 0 we have
Thus by the Arzela Ascoli theorem the sequence `n has an uniformly convergent subse-
quence, as desired.
where U and V are say compact metric spaces, viewing T as operator on different function
spaces.
Viewing T as a map from L2 (U, dµ) to L2 (V, dν) for some measure ν (note that both
spaces are separable RHilbert spaces), we know that one sufficient condition that guarantee
boundedness of T is U ×V |K|2 dµdν < ∞. It turns out that this would also imply compact-
ness of T . To see this, for each x we expand K(x, y) into the (countable) orthogonal basis
of L2 (Y, dν), which we may denote by φ1 , φ2 , . . .
∞
X
K(x, y) = Kj (x)φj (y)
j=1
note that for almost every x ∈ X the function K(x.y) is L2 (Y, dν) integrable in y and so
Z Z XZ
2
|K(x, y)| dµ(x)dν(y) = |Kj (x)|2 dµ(x)
j X
so T is compact.
76 CHAPTER 9. COMPACT OPERATORS
bounded. Note that this correction does not change yk and hence does not change the goal
of this part. Let
dk = dist(xk , ker(1 − T ))
By modifying xk by an amount inside ker(1 − T ) we may assume that dk ≤ kxk k ≤ 2dk .
Thus it suffices to show that
sup dk < ∞
k
We now prove Theorem 35. First, given any λ ∈ σ(T ) such that λ 6= 0 we show that it is
an eigenvalue with finite dimensional eigenspace. Clearly λ1 T is compact. Thus, if ker(1− λ1 T )
is trivial then by Riesz’s theorem it is also onto, therefore by the open mapping theorem
1 − λ1 T is boundedly invertible and therefore λ 6∈ σ(T ). Thus ker(1 − λ1 T ) is nontrivial and
also finite dimensional by Riesz’ theorem, and so λ is an eigenvalue with finite dim eigen
space.
Now, we will show that 0 is the only possible accumulation point, which also implies that
σ(T ) is at most countable. It suffices to show that given any > 0 there is some C = C(T, )
finite such that at most C(T, ) elements of σ(T ) would be outside [−, ]. Let λ1 , . . . , λm
be a finite collection of distinct elements of σ(T ) with |λj | > , it suffices to show that
m < OT, (1).
The idea is to let Yn be the eigenspace associated with λn and observe that for any n
it holds that Yn ∩ span{Ym , m < n} is trivial. Let Y<n = span{Ym , m < n} which is now
a strictly nested sequence. Then by Riesz’s lemma one could choose yn ∈ Yn such that
kyn k = 1 and dist(yn , Y<n ) ≥ 1/2. We will show that
kT yn − T ym k ≥ /2
for any n 6= m. Indeed, without loss of generality assume n > m, then using the fact that
yn ∈ Yn and the fact that T leaves Ym invariant we have
Let H be a Hilbert space. For a bounded operator A : H → H its Hilbert space adjoint is
an operator A∗ : H → H such that hAx, yi = hx, A∗ yi for all x, y ∈ H. We say that A is
bounded self adjoint if A = A∗ .
In this chapter we discussed several results about the spectrum of a bounded self adjoint
operator on a Hilbert space. We emphasize that in this chapter A is bounded, there is also
a notion of unbounded self adjoint operator which we will discuss in subsequent chapters.
Proof. We note that if H is the direct sum of subspaces H1 and H2 such that H1 and H2
are invariant under A then it suffices to prove the theorem for the restriction of A to each
subspace. This applies to direct sums indexed by larger index sets.
Now it is not hard to see that H could be written as an orthogonal direct sum of subspaces
of the form span(An ξ, n ≥ 0) where ξ ∈ H. (The proof uses Zorn’s lemma.) Note that
these subspaces are invariant under A, therefore it suffices to show the theorem when H =
span(An ξ, n ≥ 0) for some fixed ξ.
Now, recall from spectral caculus for bounded operators on a Banach space that if f is
analytic on a domain containing σ(A) then we could define f (A) and furthermore σ(f (A)) =
79
80 CHAPTER 10. BOUNDED SELF-ADJOINT OPERATORS
f (σ(A)). For polynomials we could do this directly using factorization into linear factors for
polynomials.
In the case of bounded self adjoint operator we will show below that σ(A) ⊂ R. (Note
that this fact also holds for unbounded case, but we will not discuss that in this section.)
Lemma 8. Let A be bounded self adjoint on a complex Hilbert space H. Then σ(A) ⊂ R.
To see this lemma, we will show that if λ ∈ C has nonzero imaginary part then λ ∈ ρ(A).
To do this, we will show that
for some c depending on λ. This would imply that λ − A is invertible using an application
of the Lax-Milgram theorem: consider the bilinear form B(x, y) = hx, (λ − A)yi, which is
nondegenerate once we proved the above estimate, thus given any z we could find y such
that hx, zi = B(x, y) for all x ∈ H, which implies that z = (λ − A)y, thus λ − A is bijective
on H and so is boundedly invertible and so λ ∈ ρ(A) as desired.
To show the above estimate, simly write λ = a + ib where a, b ∈ R and b 6= 0, then using
the self-adjoint property of A it follows that hx, (a − A)xi is a real number, therefore
To see this last claim, we first show it for real polynomial. In fact we will consider g(x) = x.
Then as we proved before
while kAk ≤ supx∈σ(A) |g(x)| using either the spectral theorem, or by elementary methods.1
The real polynomial case then follows from the spectral mapping theorem σ(g(A)) = g(σ(A))
n −n
1
Here we could easily see that kAxk2 = A2 x, x ≤ kA2 kkxk2 thus by repeating we obtain kA2 k2 ≤
kAk as desired.
10.2. PROJECTION-VALUED MEASURE AND SPECTRAL PROJECTION 81
and the fact that g(A) which is self adjoint when g is real polynomial. To allow for complex
polynomial p, simply write
thus the restriction of U to the polynomials is an isometry and the image of the polynomials
under U is clearly dense inside H by the given assumption. Thus U extends to an isometric
isomorphism from L2 (σ(A), dµ) to H.
Finally we will show that U −1 AU f (x) = xf (x) for all f ∈ L2 , note that this equality
is understood in the almost everywhere since with respect to µ, furthermore m(x) := x is
bounded in σ(A) thanks to compactness of σ(A). Thanks to Weierstrass’s theorem again it
suffices to show this equality for f being polynomials. In that case let g(x) = xf (x) also a
polynomial, we then have
R show convergence
in the weak operator norm, then it suffices to check that if f = 1Ω then
R1Ω dPλ φ1 , φ2 = h1Ω (A)φ1 , φ2 i for all φ1 , φ2 ∈ H. By definition the left hand side the same
as 1Ω dPφ1 ,φ2 = hPΩ φ1 , φ2 i which is the same as the right hand side.
continuous, point, and singular parts) using the Radon Nikodym theorem. This leads to
decomposition of Kj and also decomposition of H into the absolutely continous, singular,
and point spectrum H (p) , H (s) , and H (c) . Note that these are orthogonal subspaces and the
corresponding spectral measure of the cyclic subspace has the inherited properties. Say if
x ∈ H (c) then the specral measure of A on span(An x) is absolutely continuous with respect
to the Lebesgue measure.
Uniqueness of the spectrum Let A be bounded self adjoint on H and assume that H
be separable. Recall from prior sections that there is a decomposition of H into orthogonal
direct sum of K1 , K2 , . . . where Kj are orthogonal closed subspaces of H and each of them
is invariant under A, furthermore there is a linear isometry Uj mapping some L2 (σ(A), µj )
into Kj such that Uj−2 AUj acts on this L2 space by multiplication, in fact Uj−1 f (A)Uj g(x) =
f (x)g(x) for all bounded measureable f and g ∈ L2 (µj ). One could certainly modify dµj
multiplicatively using a bounded positive function, this would affect the isometry part of
the map U but the L2 space remains the same and the corresponding action of f (A) is still
pointwise multiplication.
Now if H has another decomposition into an orthgonal direction sum of closed subspaces
spectralSrepresentations L2 (Tj , νj ) for Lj , then upto a set of (spectral) mea-
L1 , L2 , . . . withS S
sure 0 we have Sj = Tk , namely we will show that for any k the set difference Tk \ Sj
has zero measure under νk . Note that if we assume furthermore S that these
S measures are
absoluteluy continuous with respect to the Lebesgue measure then Sj = Tk up to set of
Lebesgue measure 0.
To see this take f ∈ L2 (Tk , νk ), and let F be any bounded Borel measurable function.
Note that Sj and Tk are invariant under F (A), and f coressponds to some hf ∈ H. We now
decompose hf into hj where hj ∈ Kj , and hj in turn corresponds to fj ∈ L2 (Sj , µj ). Then
Z Z Z
|F (x)f (x)| dµk = khf k = kh1 k +kh2 k +· · · = |F (x)| |f1 | dµ1 + |F (x)|2 |f2 (x)|2 dµ2 +. . .
2 2 2 2 2 2
S
therefore if Tk \ Sj has positive µk measure we may choose F to be the characteristic
function of this set and f ≡ 1, clearly the left hand side is now positive while the right hand
side is 0, contradiction.
Spectral multiplicity
Consider the spectral representation of H, assumed separable, using the orthogonal direct
sum of K1 , . . . with Kj = L2 (Sj , µj where Sj is the support of µj .
Given λ ∈ R we define its spectral multiplicity with respect to this representation as the
number of j such that λ ∈ Sj .
Assume for simplicity that the spectrum of A is absutely continuous with respect to the
Lebesgue measure. Then it can be shown that the spectral multiplicity is independent of
the spectral representation of H; this is a theorem of Hellinger.
84 CHAPTER 10. BOUNDED SELF-ADJOINT OPERATORS
Chapter 11
and Z
Aφ = tdPt φ
We first show that σ(A) ⊂ R. To do this we will show that for every z 6∈ R the range
of A − z is closed and then we use the denseness of D to show that A − z is onto, as a by
product of the proof we will also have A − z is injective, in fact (A − z)−1 is bounded from
H to D.
We now show the spectral theorem. To do that we first show that
(i) R(z) = (A − z)−1 is a complex differentiable function of z ∈ C+ := {Im(z) > 0} in
the strong topology, namely the topology on the space of bounded operators induced by the
seminnorms Aφ, φ ∈ H. In other words, R(z)u is complex differentiable for every u ∈ H.
(ii) the adjoint of R(z) is R(z).
To see (i), first note that R(z) is a bounded operator on H for those z, let u(z) =
(A − z)−1 u for any u ∈ H. Then by simple algebra
u(z + ω) − u(z) = ωR(z)R(z + ω)u
85
86CHAPTER 11. SPECTRAL THEORY FOR UNBOUNDED SELF-ADJOINT OPERATOR
so u(z) is complex differentiable for all u, thus R(z) is complex differentiable and therefore
is analytic in the strong operator topology.
To see (ii), note that hu, R(z)vi = h(A − z)R(z)u, R(z)vi = hR(z)u, (A − z)R(z)vi =
hR(z)u, vi.
As a consequence, we know that Fu (z) := hR(z)u, ui is a complex valued analytic function
1
on the upper half plane and Fu (z) . Im(z) . Furthermore using the second property it follows
that F is rather symmetric accross the real line in the sense that F (z) = F (z).
We now show that Fu (z) has nonnegative imaginary part for any u ∈ H and z in the
upper half plane and
lim yImFu (iy) = kuk2
y→∞
To see these points, let v = R(z)u, then Fu (z) = hv, (A − z)vi = hv, Avi − z hv, vi, note
that self adjointness of A implies that hv, Avi ∈ R, therefore ImFu (z) = Im(z)kvk2 > 0 as
desired. Let z = iy where y ∈ R, then it folows that
kAR(iy)k ≤ 2
Now one could think of Fu (z) as an analytic function that maps the upper half plane
into itself, thus using Poisson integral and the Riesz representation theorem we could show
that there is some nonnegative measure µu and some linear function l(z) of z such that
Z
dµu (t)
Fu (z) = l(z) +
t−z
(this is basically Herglotz’s theorem) for all z in the upper half plane. Then combining with
the fact that yFu (iy) → kuk2 it follows that the linear part is 0 and µu (R) = kuk2 (this is
basically a result of Nevanlinna.)
87
We note that the values of Fu in the upper half plane determines its values in the lower
half plane via the relationship Fu (z) = Fu (z). Therefore the equation
Z
dµu (t)
Fu (z) =
t−z
holds for all z ∈ C \ R. Then via algebras it follows that for all u, v ∈ H
Z
dµu,v (t)
hR(z)u, vi =
t−z
where µu,v is a signed measure, basically a linear combination of µu±v , µu±iv .
1
µu,v = µu+v − µu−v + iµu+iv − iµu−iv
4
It is not hard to check the following properties
(i) µu,u = µu is a nonnegative measure with total mass kuk2 ;
(ii) µu,v is linear in u and conjugate linear in v and µu,v = µv,u
(iii) the total variation kµu,v k is controlled by kukkvk. (use Cauchy Schwartz)
Then via repeated application of the Riesz representation theorem there exists a bounded
self adjoint operator PΩ for each Borel Ω ⊂ R such that µu,v (Ω) = hPΩ u, vi for all u, v ∈ H.
We will check that PΩ indeeds define a projection valued measure, and it will then follow
that Z
1
hR(z)u, vi = d hPt u, vi
t−z
Check:
2
R (i) Clearly P= 0 since µu,v (∅) = 0 for all u, v, also PR = 1 since kuk = µU (R) =
hPt u, ui = hPR u, ui for all u ∈ H.
(ii) We will show that PΩ commutes with A. First we show that PΩ commutes with R(z).
Note that R(z1 ) and R(z2 ) commutes with each other for all z1 and z2 non real. Using
−1
R
the representation R(z)u, v >= (t − z) d hEt u, vi we have
Z Z
(t − z1 ) d hR(z2 )Pt u, vi = (t − z1 )−1 d hPt u, R(z2 )vi =
−1
= hR(z1 )u, R(z2 )vi = hR(z2 )R(z1 )u, vi = hR(z1 )R(z2 )u, vi = . . .
Z
= (t − z1 )−1 d hPt R(z2 )u, vi
Using the fact that the Cauchy transform determines uniquely the measure, it follows that
R(z2 )PΩ = PΩ R(z2 ) for any Borel Ω.
Let u ∈ H, then AR(z)u = u + zR(z)u, thus for any Ω we have
Fredholm determinant
12.1 Motivation
Fredholm determinant started in Fredholm’s investigation of the integral equation
(1 + K)u = f
R
where K is the integral operator Kf (y) = Y K(x, y)f (y)dy maping functions on Y to
functions on X. (Here X, Y are compact metric spaces.) We know that K is compact
from L2 (Y ) to L2 (X) if the kernel K is square integrable in X × Y . The setting originally
considered is for continuous functions, for which we have the following results:
(i) If K is a continuous function of x and y then K is compact from L1 (Y ) to C(X).
(ii) If K is a continuous function of x in the L1 norm wrt to y, i.e. kK(x, .)kL1y is
continuous wrt x, then K is compact from C(Y ) to C(X).
Proof of these facts uses the Arzela Ascoli criteria: S haudorff compact, functions equicon-
tinuous and pointwise uniformly bounded, then the family is precompact in the sup norm
on S.
For simplicity, assume K : [0, 1]2 → C is continuous and f ∈ C[0, 1]. Note that this is
not a Hilbert space. By Riesz’s theorem the integral equation
Z 1
u(x) + K(x, y)u(y)dy = f (x)
0
is solvable on C[0, 1] (i.e. given f ∈ C[0, 1] one could find u ∈ C[0, 1]) iff the integral operator
K is injective on this space, or equivalently its range is the whole space. It can be shown
that a vector is in the range of K iff it is orthogonal to the null space of K ∗ which has kernel
K ∗ (x, y) = K(y, x).
Fredholm investigated the above equation by discretizing the equation and appeal to
linear algebra, and then taking a limit at the end. This give rises to a number called the
Fredholm determinant of 1 + K (we simply say the Fredholm determinant for K), which
determines whether the given integral equation is solvable or not. The determinant concept
89
90 CHAPTER 12. FREDHOLM DETERMINANT
has been extended to other settings, most commonly for K being a trace class operator on
some separable Hilbert space.
We first discuss Fredholm’s orginal approach for C[0, 1] (modulo some simplifications),
and then we’ll discuss the Hilbert space theory later.
1 d m
cm = ( ) D(h)|h=0
m! dh
We then use the product rule, which say that if C = det(C1 , P . . . , Cn ) is the determinant of
d d
a matrix with columns C1 , . . . , Cn then by multilinearity dh C = k det(C1 , . . . , dh Ck , . . . , Cn ).
In our case each column is linear in h, and Cj (0) = (0, . . . , 1, . . . , 0) the unit is in the jth
position. Therefore
h2 X
X Kii Kij
D(h) = 1 + h Kjj + det + ...
2 Kji Kjj
j i,j
x1 , . . . , xk
For convenience let K denote the determinant of the matrix K(xi , yj ), then
y1 , . . . , yk
letting h = 1/n and send n → ∞ we obtain
∞ Z Z
X 1 x1 , . . . , xk
D= ... K dx1 . . . dxk
k! x1 , . . . , xk
k=0
This is called the Fredholm determinant of the integral operator K. Note that this is a
complex number.
12.2. FREDHOLM’S APPROACH FOR INTEGRAL OPERATORS 91
This follows from the fact that the volume of the parallelopipde is at most the product of
the side lengths.
Now, the given assumption on K implies that supx,y |K(x, y)| ≤ M for some finite M .
So by Hadamard inequality we have
x1 , . . . , xk
K ≤ (M k 1/2 )k (12.1)
y1 , . . . , yk
Proof. For convenience, let F be the n × n matrix with entries F (xi , xj ) and G be the n × n
matrix with entries G(xi , xj ). It is clear that
where Mk is the matrix whose first k − 1 rows are the same as G, and the kth row is the
same as F − G, and the last n − k rows are the same as F .
Apply Hadamard inequality we obtain the first desired estimate.
We now discuss how to solve the integral equation using Fredholm’s determinant. The
idea is to solve it at the discrete level and then send h → 0 via n → ∞. Via this heuristic
we obtain
(I + K)(I + L) = (I + L)(I + K) = I
where Z
Lf = L(x, y)f (y)dy
X 1 Z Z
−1 x, x1 , . . . , xk
L(x, y) = −D ... K dξ1 . . . dξk
k! y, y1 , . . . , yk
k≥0
Z
L(x, y) + K(x, y) + L(x, z)K(z, y)dz = 0
Then L = − D1 R.
x, x1 , . . . , xk
Now, computing the determinant K using its first row, we obtain
y, y1 , . . . , yk
x, x1 , . . . , xk x1 , x2 , . . . , xk
K = K(x, y)K +
y, y1 , . . . , yk y1 , y2 , . . . , yk
k
X
k x1 , x 2 , ..., xk
+ (−1) K(x, yj )
y, y1 , . . . , (yj ), . . . yk
j=1
for every x, y. If R 6≡ 0 then one could find one y such that g(.) = R(., y) is not the zero
function (note that it is continuous), and it satisfies g + Kg = 0, therefore 1 + K is not
injective.
It is however possible that R ≡ 0. If so, consider the following functions
X zk Z Z
x1 , . . . , xk
D(z) = ... K dx1 . . . dxk
k! x1 , . . . , xk
k≥0
X z k+1 Z Z
x, x1 , . . . , xk
R(x, y, z) := ... K dξ1 . . . dξk
k! y, y1 , . . . , yk
k≥0
One could think of D(z) and R(x, y, z) as the version of D and R with zK instead of K.
It is clear that D and R are entire functions. Since D(1) = 0 the entire function D has a
zero of finite order n ≥ 1 at z = 1. By algebraic manipulation we have
Z
R(x, x, z)dx = zD0 (z)
therefore there is some (x, y) such that R(x, y, z) can not vanish at z = 1 with order more
than n − 1. Then for some 1 ≤ ` < n it holds that
where g(x, y) 6≡ 0, and g is continuous in x, y (to see this note that g is the uniform limit of
a Rsequence of continuous functions on [0, 1]2 ). We recall that R(x, y, z) = zK(x, y)D(z) −
z K(x, x1 )R(x1 , y, z)dx1 thus by dividing everything by (z − 1)` and then send z → 1 we
obtain Z
g(x, y) = − K(x, x1 )g(x1 , y)dx1
Theorem 41. Assume that K is Holder c-continuous where c > 1/2. Then the nonzero
eigenvalues
P of K on C[0, 1] (only countablity many of them since K is compact) satisfies
j |λj | < ∞ and for every z ∈ C we have
Y
D(z) = (1 + zλj )
j
Proof. We recall Hadamard’s factorization theorem (or may be a consequence of this theo-
rem): Let f be an entire function such that for some finite positive C1 , C2 and ρ ∈ [0, 1) it
holds for every complex number z that
Assume that f (0) 6= 0. Then f has at most a countable number of zeros, furthermore
X 1
<∞
|z|
z: f (z)=0
We plan to use this theorem to show the first part of the theorem. We note that the
second part of the theorem, i.e. the trace formula would then follows from
Z
K(x, x)dx = D0 (z)|z=0
Q
(which is part of the definiton of D) and the absolute convergence of the product j (1+zλj )
(viewed as an infinite power series for z).
First, we will show that for z 6= 0 it holds that: D(z) = 0 if and only if −1/z is an
eigenvalue of K. This is simply a consequence of our last theorem applied to D = det(1+zK).
Now, we will show that if K is Holder c-continuous then
2
|D(z)| . exp(O(|z| 1+2c )
To see this, using Stirling’s formula it suffices to show that, for some C finite,
X C n |z|2n/(1+2c)
|D(z)| .
n≥0
nn
12.3. FREDHOLM DETERMINANT FOR OPERATORS ON HILBERT SPACES 95
Using the definition of D it suffices to show that given any x1 , . . . , xk , y1 , . . . , yk it holds that
for some C > 0 finite
x1 , x2 , . . . , xk
|K | ≤ (Ck 1/2 )k k −c
y1 , y2 , . . . , yk
Note that incomparision with (12.1) we gain a factor of k −c . To see this it suffices to show
that k−1
x1 , x2 , . . . , xk 1/2
Y
|K | ≤ (Ck )k |yj+1 − yj |c
y1 , y2 , . . . , yk
j=1
(note that the constant C may be different in different display). Indeed, wlog weQ
may assume
y1 ≤ y2 · · · ≤ yk , then by the geometric arirthmetic mean inequality we have k−1 j=1 |yj+1 −
1 k−1
yj | ≤ ( k−1 ) , as desired.
Now, note that if c1 , . . . , ck are k column vectors (k × 1) then
Definition: We say that T is a trace class operator if the sum of its singular values is
finite (counting with multiplicity). In that case the trace norm of T is defined to be this
sum, denoted by kT ktr .
Properties: T and T ∗ has the same trace norm, and if T is trace class then so is T B
and BT where B is any bounded operator on H, and k.ktr satisfies the triangle inequality.
where the sup is taken over all (fn ) and (en ) orthonormal bases of H.
To show this characterization we will first derive the polar factorization of T : namely
there is a partial unitary operator U such that T = U A, here unitary of U simply means
that U ∗ U when acting on the range of A is the identity operator. This should be thought of
as the operator analogue of the usual polar factorization of a complex number.
To define U , note that kAuk = kT uk for all u ∈ H, therefore we may define an isometry
U from range(A) to range(T ) by mapping Au to T u.
One then extends U to H by letting U to be zero on the orthogonal complement of
range(A). It follows immediately that U ∗ H ⊂ range(A): one simply notice that for every z
in the othogonal complement of range(A) and z 0 ∈ H it holds that hz, U ∗ z 0 i = hU z, z 0 i = 0.
Now let z in the closure of the range of A. We want to show that (U ∗ U − 1)z = 0, which
is the desired local unitary property of U . Using the isometric property of U on range(A),
for every z 0 in the closure of the range of A we have
hz 0 , zi = hU z 0 , U zi = hz 0 , U ∗ U zi
for any pair of orthogonal bases (fn ) and (en ). Let sn be the singular values of T , namely
AFn = sn Fn . Then expand fn into (Fn ) we have
X
fn = hfn , Fk i Fk
k
(it will be clear from the proof that the double sum is absolutely summable)
X X X
≤ |sk |( | hfn , Fk i |2 )1/2 ( | hGk , en i |2 )1/2
k n n
X X
≤ |sk |kFk kkGk k = |sk | = kT ktr
k k
where (fn ) is any orthonormal basis of H. This definition is independent of the choice of the
basis. Let (gn ) be another orthonormal basis, then by expanding fn into this new basis we
have X XX
hT fn , fn i = hfn , gk i hT gk , fn i
n n k
it is not hard to see that this double sum is abs convergence (using Cauchy Schwartz). Then
by Fubini X XX
hT fn , fn i = ( hfn , gk i hT gk , fn i)
n k n
X
= hT gk , gk i
k
By definition it is clear that |T race(T )| ≤ kT ktr and the sum defining the trace is
absolutely convergent.
Now, Lidskii’s trace formula says that
98 CHAPTER 12. FREDHOLM DETERMINANT
Note that by definition we have kTn − T k → 0 (in operator norm) and T race(Tn ) →
T race(T ). Since the spectral radius of T is 0 it is clear that the spectral radius σ(Tn )
of Tn converges to 0 too. In particular given any bounded set of the complex plane we may
choose n large such that this bounded set is contained inside the ball of radius 1/σ(Tn ).
Furthermore one could also show that sj (Tn ) ≤ sj (T ) if the singular values of T and Tn are
ordered in decreasing order.
Now, it is not hard to see that
Dn0 (λ) X α
=
Dn (λ) α∈Λ 1 + λα
n
X
= T race(Tn ) + O (|λ|σ(Tn ))k−1 kTn kT r
k≥2
Thanks to convexity of log, one could show that this inequality is a consequence of the fact
that
Y N
Y
|α| ≤ sj (Tn )
α∈Λn j=1
thus using the previous argument it suffices to show that T ∗ leaves K2 invariant and the
only eigenvalue of T ∗ on K2 is 0 (it is clear that T2 is also compact and trace class). These
properties could be easily checked.
Determinant We now discuss det(1 + T ) for trace class operators T on a separable
Hilbert space H.
Define an inner product on H k by
One could check that if T is finite rank then this determinant is the same as what we
would obtain from linear algebra, and det(1 + .) is locally Lipschitz wrt to the trace class
norm. Approximating T by a sequence of finite rank operators (for which det(1 + T ) could
be defined using standard linear algebra), we could show that det(1 + T ) is the limit of the
corresponding determinants, and thus this limit is independent of the choice of the sequence.
Now, one could use the polar decomposition T = U A and approximate A by projections into
finite dimensional subspaces of H spanned by eigenfunctions of A. By this approximation
scheme it can be shown that Y
det(1 + T ) = (1 + λj )
where λj are the eigenvalues of T and
and the definition of the determinant coincides with the previous section.