Lectures in Functional Analysis-Roman Vershynin PDF
Lectures in Functional Analysis-Roman Vershynin PDF
Roman Vershynin
Department of Mathematics, University of Michigan, 530 Church
St., Ann Arbor, MI 48109, U.S.A.
E-mail address: romanv@umich.edu
Preface
These notes are for a one-semester graduate course in Functional Analysis,
which is based on measure theory. The notes correspond to the course Real Analysis
II, which the author taught at University of Michigan in the Fall 2010. The course
consists of about 40 lectures 50 minutes each.
The student is assumed to be familiar with measure theory (both Lebesgue and
abstract), have a good command of basic real analysis (epsilon-delta) and abstract
linear algebra (linear spaces and transformations).
The course develops the theory of Banach and Hilbert spaces and bounded
linear operators. Main principles of are covered in depth, which include Hahn-
Banach theorem, open mapping theorem, closed graph theorem, principle of uni-
form boundedness, and Banach-Alaoglu theorem. Fourier series are developed for
general orthogonal systems in Hilbert spaces. Compact operators and basics of
Fredholm theory are covered.
Spectral theory for bounded operators is studied in the second half of the course.
This includes the spectral theory for compact self-adjoint operators, functional
calculus and basic spectral theory of general (non-compact) operators, although
the latter needs to be expanded a bit.
Topics not covered include: Krein-Milman theorem (although this can be done
with one extra lecture), unbounded linear operators, and Fourier transform. Most
applications to ODE and PDE are not covered, however the integral operators serve
as a main example of operators in this course.
The material has been compiled from several textbooks, including Eidelman,
Milman and Tsolomitis “Functional Analysis”, Kirillov and Gvishiani “Theorems
and problemsin functional analysis”, Reed and Simon “Methods of modern mathe-
matical physics. I. Functional analysis”, V. Kadets “A course in functional analy-
sis” (Russian), and P. Knyazev, “Functional analysis”. Minor borrowings are made
from Yoshida “Functional analysis”, Rudin “Functional analysis”, and Conway “A
course in functional analysis”. For some topics not covered, one may try R. Zimmer
“Essential results of functional analysis”.
Acknowledgement. The author is grateful to his students in the Math 602
course Real Analysis II, Winter 2010, who suggested numerous corrections for these
notes. Special thanks are to Matthew Masarik for his numerous thoughtful remarks,
corrections, and suggestions, which improved the presentation of this material.
Contents
Preface iv
Chapter 1. Banach and Hilbert spaces 1
1.1. Linear spaces and linear operators 1
1.2. Normed spaces 7
1.3. Banach spaces 15
1.4. Inner product spaces 19
1.5. Hilbert spaces 25
1.6. Fourier series 28
Chapter 2. Bounded linear operators 39
2.1. Bounded linear functionals 39
2.2. Representation theorems for linear functionals 42
2.3. Hahn-Banach theorem 47
2.4. Bounded linear operators 56
Chapter 3. Main principles of functional analysis 69
3.1. Open mapping theorem 69
3.2. Closed graph theorem 74
3.3. Principle of uniform boundedness 76
3.4. Compact sets in Banach spaces 83
3.5. Weak topology 87
3.6. Weak topology. Banach-Alauglu’s theorem 91
Chapter 4. Compact operators. Elements of spectral theory 94
4.1. Compact operators 94
4.2. Fredholm theory 98
4.3. Spectrum of a bounded linear operator 101
4.4. Properties of spectrum. Spectrum of compact operators. 103
Chapter 5. Self-adjoint operators on Hilbert space 110
5.1. Spectrum of self-adjoint operators 110
5.2. Spectral theorem for compact self-adjoint operators 113
5.3. Positive operators. Continuous functional calculus 116
5.4. Borel functional calculus. Spectral theorem for self-adjoint operators 122
Bibliography 126
Index 127
v
CHAPTER 1
This is clearly a metric, so a function space becomes not only a linear vector space
but also a metric space. Such spaces will be called normed spaces later. Another
natural choice of a distance would be the
»b
}f g}1 : |f pxq gpxq| dx
a
Lec. 2: 09/10/10
1.1.4. Hamel basis. As we know, every finite dimensional linear vector space
E has a basis tx1 , . . . , xn u. A basis is a maximal linearly independent subset of
vectors in E. The number n of basis elements is called the dimension of E; this
number is independent of the choice of the basis. Every vector x P E can be
uniquely expressed as a linear combination of the basis elements:
¸
n
(1.1) x ak xi , for some ak P R.
k 1
The notion of basis can be generalized to arbitrary, possibly infinite dimensional
linear vector spaces E.
pq pq
1For consistency of these inclusions, we restrict the functions in P x , P x and F onto
n
ra, bs.
1.1. LINEAR SPACES AND LINEAR OPERATORS 3
1. dimpRn q n, dimpCn q n.
2. dimpPn pxqq n 1, the monomials t1, x, x2 , . . . , xn u form a basis.
3. dimpP pxqq 8, the monomials t1, x, x2 , . . .u form a Hamel basis.
4. dimpc00 q 8, the coordinate vectors p0, . . . , 0, 1, 0, . . .q form a Hamel basis.
Remark 1.1.9. Unfortunately, the notion of Hamel basis is too strong. Except
in spaces P pxq and c00 (which are isomorphic - why?) no explicit constructions are
known in any other infinite dimensional vector space. It would be great to have
a construction of a Hamel basis in C r0, 1s, for example. However, Hamel bases
usually have to be uncountable; see a later exercise.
1.1.5. Quotient spaces. The notion of quotient space allows one easily to
collapse some directions in linear vector spaces. One reason for doing this is when
one has unimportant directions and would likes to neglect them; see the construction
of L1 below.
Definition 1.1.10 (Quotient space). Let E1 be a subspace of a linear vector
space E. Consider an equivalence relation on E defined as
xy if xy P E1 .
The quotient space E {E1 is then defined as the set of equivalence classes (cosets)
rxs for all x P E.
The quotient space is a linear space, with operations defined as
rxs rys : rx y s, arxs : raxs for x, y P E, a P R.
The dimension of the quotient space is called the codimension of E1 , thus
codimpE1 q : dimpE {E1 q.
Exercise 1.1.11. Prove that the operations above are well defined,
and that quotient space is indeed a linear space.
This way, the elements of L1 are, strictly speaking, not functions but classes of
equivalences.3 But in practice, one thinks of an f P L1 as a function, keeping in
mind that functions that coincide µ-almost everywhere are “the same”.
Example 1.1.14 (Space L8 ). A similar procedure is used to define the space
of essentially bounded functions L8 L8 pΩ, Σ, µq. A real valued function f on Ω
is called essentially bounded if there exists a bounded function g on Ω such that
f g µ-almost everywhere. Similar to the previous example, we consider the linear
vector space
E : tall essentially bounded functions f on pΩ, Σ, µqu
and the subspace we would like to neglect:
E1 : tall functions f 0 µ-almost everywhereu
Then we define
L8 L8 pΩ, Σ, µq : E {E1 .
Example 1.1.15. As we know, the space c0 of sequences converging to zero is
a subspace of the space c of all convergent sequences. Let us observe that
codimpc0 q 1.
Indeed, every sequence x P c can be uniquely represented as
x a1 z for some a P R, z P c0
where 1 p1, 1, . . .q. (How do we choose the value of a?). Hence
rxs ar1s rzs ar1s.
It follows that every element rxs P c{c0 is a constant multiple of the element r1s.
Therefore, dimpc{c0 q 1 as claimed.
This example shows that the space c0 makes up almost the whole space c,
except for one dimension given by the constant sequences. This explains why the
space c is rarely used in practice; one prefers to work with c0 which is almost the
same as c but has the advantage that we know the limits of all sequences there
(zero).
1.1.6. Linear operators. This is a quick review of the classical linear algebra
concept.
Definition 1.1.16 (Linear operator). A map T : E Ñ F between two linear
vector spaces E and F is called a linear operator if it preserves the operations of
addition of vectors and multiplication by scalars, i.e.
T pax by q aT pxq bT py q for all x, y P E, a, b P R.
4
The kernel and image of T is defined respectively as
kerpT q tx P E : T x 0u; ImpT q tT x : x P E u.
3Even more strictly speaking, the representative functions f in L may take infinite values,
1
too. However, every integrable function is finite a.e. So every such function is equivalent to a
function that is finite everywhere.
pq
4One usually writes T x instead of T x
1.1. LINEAR SPACES AND LINEAR OPERATORS 6
2. The spaces c and c0 are normed spaces, with the same sup-norm as in
(1.4).
3. The space of summable sequences `1 is a normed space, with the norm
defined as
8̧
}x}1 : |xi |.
i 1
1.2. NORMED SPACES 8
Here the infimum is taken over all g P L8 that are equal to f µ-almost
everywhere. Note that `8 is a partial case of the space L8 pΩ, Σ, µq
where Ω N and µ is the counting measure on N.
Definition 1.2.6 (Convex functions and sets). Let E be a linear vector space.
A function f : E Ñ R is convex if for all x, y P E, λ P r0, 1s one has
f pλx p1 λqyq ¤ λf pxq p1 λqf pyq.
A set K E is convex if for all x, y P K, λ P r0, 1s one has
λx p1 λqy P K.
A geometric meaning of convexity is the following. A function f is convex on
E if its graph restricted to any interval rx, y s E lies below the interval joining
the points px, f pxqq and py, f py qq; see the picture. A set K is convex if, together
with any two points x, y P K, it contains the interval rx, y s.
1. If the function x Ñ }x} is convex, then the triangle inequality is satisfied, and
} } is a norm on E.
2. If the sublevel set tx P X : }x} ¤ 1u is convex, then } } is a norm on E.
Proof. 1. Convexity ensures that for every x, yP E, λ P r0, 1s we have
}λx p1 λqyq} ¤ λ}x} p1 λq}y}.
Triangle inequality follows from this for λ 1{2.
2. This statement is less trivial, and can not be obtained from the first one
directly. Indeed, while it is true that the sublevel sets of a convex functions are
convex sets, the converse statement may fail (construct an example!)
The assumption states that, for u, v P E we have:
(1.5) if }u} ¤ 1, }v } ¤ 1, λ P r0, 1s, }λu p1 λqv} ¤ 1.
then
Let x, y P E; we want to show that }x y } ¤ }x} }y }. This is equivalent to
u
x
}x} , v }yy} , λ }x}}x}}y} .
This completes the proof.
Lec. 4: 09/15/2010
1.2.3. Spaces Lp . Minkowski inequality. We have already come across
the spaces L1 and `1 . They are partial cases of a big family of spaces Lp and `p
which we will study now.
Consider a measure space pΩ, Σ, µq and an exponent p P r1, 8q. We define
the space of p-integrable functions Lp Lp pΩ, Σ, µq as the set of all measurable
functions f : Ω Ñ R such that
»
|f pxq|p dµ 8.
Ω
Proof. Norm axioms (i) and (ii) are straightforward. Axiom (iii), triangle
inequality, will follow from Proposition 1.2.8. To this end, it suffices to check that
the sublevel set
Bp : tf P Lp : }f }p ¤ 1u
is a convex set. To prove this, let us fix f, g P Bp and λ P r0, 1s. Since the function
z ÞÑ |z |p is convex on R for p ¥ 1, we have a pointwise inequality
λf t p q p1 λqgptqp ¤ λ|f ptq|p p1 λq|gptq|p .
Integrating both sides of this inequality implies
» » »
p q p1 λqgptqp dµ ¤ λ |f ptq|p dµ p1 λq |gptq|p dµ ¤ λ p1 λq 1.
λf t
Ω Ω Ω
We have showed that }λf p1 λqg }p ¤ 1, hence λf p1 λqg P Bp . Therefore,
the sublevel set Bp is convex. The proof is finished by Proposition 1.2.8.
Writing out the triangle inequality }f g }p ¤ }f }p }g}p in analytic form, we
obtain the classical Minkowski inequality:
Theorem 1.2.11 (Minkowski inequality in Lp ). Let p P r1, 8q. Then, for every
two functions f, g P Lp pΩ, Σ, µq one has
» 1{p » 1{p » 1{p
|f ptq g ptq|p dµ ¤ |f ptq|p dµ |gptq|p dµ .
Ω Ω Ω
1.2.4. Spaces `p and `np . An important partial case of the space Lp pΩ, Σ, µq
is the space `p obtained by choosing Ω N and µ to be the counting measure on
N. Equivalently, for p P r1, 8q, the space of p-summable sequences `p is defined to
consist of sequences x pxi q8
i1 for which
8̧
|xi |p 8.
i 1
We turn `p into a normed space with the norm
8̧ 1{p
}x}p : |xi |p .
i 1
Writing down Minkowski inequality for this specific measure space, we obtain:
Theorem 1.2.12 (Minkowski inequality in `p ). Let p P r1, 8q. Then, for every
two sequences x, y P `p one has
8̧ 1{p 8̧ 1{p 8̧ 1{p
|xi yi |p ¤ |xi |p |yi |p .
i 1
i 1
i 1
A remarkable family of finite-dimensional spaces Lp pΩ, Σ, µq is formed by con-
sidering Ω to be a finite set, say Ω t1, . . . , nu and µ to be the counting measure on
Ω. The resulting space is called `np . The functions in `np can be obviously identified
with vectors in Rn . Thus `np pRn , } }p q with the norm
¸
n 1{p
}x}p : |xi |p .
i 1
1.2. NORMED SPACES 12
}x}p Ñ }x}8 as p0 ¤ p Ñ 8.
2. Consider the space L8 L8 pΩ, Σ, µq with finite total measure µpΩq.
Show that if f P L8 then
}f }p Ñ }f }8 as p Ñ 8.
1.2.5. Subspaces of normed spaces.
Definition 1.2.14 (Subspace). Let X be a normed space. A subspace Y of X
is a linear subspace equipped with the norm induced from that of X.
This concept should be familiar from topology, where a subspace is a subset of
a topological space with the induced topology.
Example 1.2.15. 1. The space of polynomials P pxq is a dense subspace of
C r0, 1s. This is the statement of Weierstrass approximation theorem.
2. The set of all continuous functions C r0, 1s forms a dense subspace of L1 r0, 1s. (of
course, both spaces are considered in the L1 norm!) This follows from a theorem
in measure theory that states that an integrable function can be approximated
by a continuous function (why?)
Exercise 1.2.16. 1. Show that the set of convergent sequences c and
the set of sequences converging to zero c0 are closed subspaces of `8 .
2. For all p P r1, 8q, show that the set of p-summable sequences `p is a
closed subspace of `8 but is a dense subspace of c0 .
Show that } }K is a norm on Rn , and the unit ball of this normed space
is K.
normed space is deficient; many results in analysis can not be obtained just based
on the norm axioms. An additional axiom is needed, which is completeness.
Recall that a metric space X is called complete if every Cauchy sequence in X
converges to a point in X. For example, R is a complete metric space while Q is
not.
Specializing to normed spaces X, recall that a sequence pxi q8
i1 in X is Cauchy
if
}xn xm } Ñ 0 as n, m Ñ 8,
i.e. for every ε ¡ 0 there exists N N pεq such that
}xn xm } ε for all n, m ¡ N.
Definition 1.3.1 (Banach space). A complete normed space X is called a
Banach space.
Exercise 1.3.2. [Subspaces, quotients of Banach spaces] Let X be a
Banach space and Y be a (linear) subspace of X. Show that:
1. Y is a Banach space if and only if Y is closed.
2. If Y is closed8, then X {Y is a Banach space.
Many classical spaces are Banach spaces.
Theorem 1.3.3. For a compact topological space K, C pK q is a Banach space.
Proof. Most proofs of completeness work out by reducing the problem to the
completeness of R. To this end, consider a Cauchy sequence pfn q in C pK q, that is
(1.7) }fn fm }8 Ñ 0, n, m Ñ 8.
Therefore, for every t P K, we have |fn ptq fm ptq| Ñ 0. In other words, pfn ptqq
is a Cauchy sequence in R for every t. By completeness of R, this sequence has a
limit which we call f ptq. We have constructed a function f ptq such that fn Ñ f
pointwise.
We now claim that fn Ñ f uniformly, i.e. }fn f }8 Ñ 0. This would
complete the proof, since the limit of a uniformly convergent sequence of continuous
functions on the compact space K is a continuous function (as we know from the
undergraduate analysis). Let us prove our claim then. By (1.7), for every ε there
exists N N pεq such that
|fn ptq fm ptq| ε for all n, m ¡ N, t P K.
Letting m Ñ 8 in this inequality (while keeping everything else fixed), we conclude
that
|fn ptq f ptq| ε for all n, m ¡ N, t P K.
This means that }fn f }8 Ñ 8, which is what we wanted.
Exercise 1.3.4. [Banach spaces `8 , c0 , L8 ]
1. Show that `8 and L8 are Banach spaces. (Hint: modify the proof of
Theorem 1.3.3.)
2. Show that c0 is a Banach space (Hint: use Exercise 1.2.16.)
{
8Recall that Y has to be closed in order for X Y to be well defined.
1.3. BANACH SPACES 16
Theorem 1.3.8. For every p P r1, 8q, the space Lp Lp pΩ, Σ, µq is a Banach
space.
Proof. Let functions pfn q in Lp form the terms of an absolutely convergence
series, i.e.
8̧
}fk }p : M 8.
k 1
By
° the completeness criterion, Theorem 1.3.6, it suffices to show that the series
k fk converges in Lp . °n
Case 1: all fk ¥ 0 pointwise. The partial sums k1 fk form a pointwise non-
°8
decreasing sequence of functions. Denote the pointwise limit by k1 fk ; it may
be infinite at some points.
The triangle inequality (which is Minkowski’s inequality) implies that the par-
tial sums are bounded:
¸n ¸
n
fk ¤ }fk }p ¤ M,
k 1
p
k 1
which in other words is » ¸
n p
fk dµ ¤ M p .
Ω
k 1
1.3. BANACH SPACES 18
We
°napplyLebesgue monotone convergence theorem for the sequence of functions
p
k 1 f k and get
» ¸
n p » 8̧ p
fk dµ Ñ fk dµ.
Ω
k 1 Ω k 1
The right hand ° side exists and is finite since the left hand side is bounded by M p
8
for all n. Thus k1 fk P Lp .
°8
It remains to check that the series k1 fk P Lp converges not only pointwise
but also in Lp . By the pointwise convergence and positivity, the tails of this series
8̧
rn : fk
k n 1
Then turn this space into a normed space Xp by taking quotient over the
kernel of the semi-norm (see Exercise 1.2.23).
The concept of completion suggests an alternative construction of the space
Lp ra, bs, p P r1, 8q. From measure theory we know that the set of continuous
functions is a dense subset of Lp ra, bs (why?). Furthermore, Lp ra, bs is a com-
plete normed space. Therefore (by the uniqueness of completion), Lp ra, bs is the
completion of the space C ra, bs in the } }p norm.
This is an alternative definition of Lp ra, bs. This gives also an alternative con-
struction of Lebesgue integral. Indeed, on for continuous functions, Riemann and
Lebesgue integrals coincide. Therefore, the space C ra, bs with } }1 can be con-
structed using only Riemann integral. But its completion gives rise to Lebesgue
integral.
Remark 1.4.2. The inner product is (congjugate) linear in the second argu-
ment:
xx, ay bzy axx, yy bxx, zy.
This follows from axioms (ii) and (iii) of the inner product.
Definition 1.4.3 (Orthogonality). If xx, y y 0 we say that vectors x and y
are orthogonal and write x K y.
Example 1.4.4. The canonical example of a finite-dimensional Hilbert space
is the Euclidean space Cn equipped with the inner product
¸
n
xx, yy xk yk .
k 1
Remark 1.4.8 (Angle between vectors). The concept of inner product makes
it possible to define the angle between two vectors x, y in an inner product space
X. Recall that in Euclidean space Rn , the inner product can be computed by the
formula
xx, yy }x}}y} cos θpx, yq
where θpx, y q is the angle between x and y. Therefore, in a general inner product
space X, it makes sense to define the angle between x, y by
cos θpx, y q
xx, yy .
}x}}y}
Cauchy-Schwarz inequality guarantees that the right hand side lies in r1, 1s, so
the angle exists. Nevertheless, the concept of angle is rarely used; one prefers to
work with inner product directly.
Lec.8: 09/24/10
The left
³ hand side of Cauchy-Schwarz inequality can be replaced by the larger
quantity |f g | dµ. (This can be seen by applying Cauchy-Schwarz inequality for
|f |, |g|.) Thus Cauchy-Schwarz inequality can be written as
}f g}1 ¤ }f }2 }g}2 .
Cauchy-Schwarz inequality is a partial case of the more general Hölder’s in-
equality:
Corollary 1.4.11 (Hölder’s inequality for functions). Let p, q P p1, 8q be
adjoint, i.e. p1 1q 1. Then for every f P Lp and g P Lq one has
» » 1{p » 1{q
f ḡ dµ ¤ |f |
p
dµ |g|q dµ .
However, Corollary 1.4.12 on the scale of Lp (on finite measure spaces) does
not hold for `p , because the counting measure on N is not finite. In fact, the scale
is completely reversed in this case: `1 is the smallest space, `8 is the largest, and
the other `p , p P r1, 8q lie in between:
Corollary 1.4.15 (Scale of `p spaces). Let 1 ¤ r ¤ s ¤ 8. Then
}x}s ¤ }x}r for all x P `r .
In particular, we have the inclusion
`r `s .
Exercise 1.4.16. Prove Corollary 1.4.15. You don’t have to use Hölder’s
inequality.
This is clearly an inner product. One way to see this is to identify Mm,n with Cmn
by concatenating the rows of a matrix A P Mm,n into a long vector in Cmn . Then
the canonical inner product in Cmn is the same as the right hand side of (1.11).
The norm defined by the inner product on Mm,n is called Hilbert-Schmidt or
Frobenius norm of matrices:
¸
m ¸
n 1{2
(1.12) }A}HS xA, Ay1{2 |aij |2
i 1j 1
Note³ some similarity between the forms of the inner product in L2 , which is
xf, gy f ḡ dµ and in Mm,n , which is xA, B y : trpAB q – the integral is replaced
by the trace, functions by matrices, complex conjugation by transposition, and
product of functions by product of matrices.
Therefore, the space L2 pΩ, Σ, Pq consists of all random variables X with finite
second moment:
}X }2 pEX 2 q1{2 8.
Exercise 1.4.17. Show that X P L2 pΩ, Σ, Pq if and only if X has finite
variance:
VarpX q EpX EX q2 EX 2 pEX q2 8.
The concepts of covariance and correlation coefficient have some geometric
meaning, too. Consider two random variables X and Y , and for simplicity assume
that they have mean zero, i.e. EX EY 0. Then the covariance of X and Y is
nothing else than the inner product in L2 :
covpX, Y q : EpX EX qpY EY q EXY xX, Y y.
Similarly, the correlation coefficient between X and Y is
covpX, Y q
corrpX, Y q :
VarpX q1{2 VarpY q1{2
pEX 2 qEXY
1{2 pEY 2 q1{2
}XxX, Yy
} }Y } .
2 2
Hence the correlation coefficient is nothing else as the cosine of the angle between
random variables X and Y considered as vectors in L2 (see Remark 1.4.8). This
demonstrates the geometric meaning of correlation – the more random variables X
and Y are correlated, the less the angle between them, and vice versa. Lec.9: 09/27
The parallelogram law characterizes inner product spaces. First recall that in
inner product spaces, the inner product determines the norm (}x} xx, xy1{2 ).
Vice versa, the inner product is uniquely determined by the norm, and it can be
reconstructed through the polarization identity:
Proposition 1.4.19 (Polarization identity). Let X be an inner product space.
Then for every x, y P X one has
(1.14) xx, yy 14 }x y }2 }x y }2 i}x iy }2 i}x y }2 .
The proof of this result is deferred to the exercises for this section.
It follows from Theorem 1.4.21 that being an inner product space is a “local”
property, since checking the parallelogram law involves just two (arbitrary) vectors.
In particular, if all two-dimensional linear subspaces of a normed space X are inner-
product spaces (with respect to some inner product, possibly different for each
subspace), then X is an inner product space (and there the inner product on all
subspaces is actually the same, induced from X!)
Exercise 1.4.25. Show that C pK q, c0 , Lp r0, 1s, `p for p P r1, 8s, p 2, are
not inner product spaces. (More accurately, it is not possible to define
an inner product on those spaces which would agree with their norms).
Use Theorem 1.4.21.
To bound }yn ym }, we use parallelogram law. We apply it for the parallelogram
with vertices x, yn , ym (and whose fourth vertex is determined by these three, see
the picture.)
}yn ym }2 4}x
1
2
pyn ym q}2 2}x yn }2 2}x ym }2 .
In the proof of part (i) of Theorem 1.5.5, we used convexity rather than linearity
of Y . (Indeed, we needed that together with two points yn , ym P Y their midpoint
2 pyn ym q is contained in Y ). Therefore, our argument implies the following more
1
general result:
Theorem 1.5.6 (Hilbert’s projection theorem). Given a closed convex set Y
in a Hilbert space X and a point x P X, there exists a unique closest point y P Y .
The map that takes x into the closest point y is called a projection onto convex
set Y and is abbreviated POCS. This map appears in several applied fields.
The orthogonality principle immediately implies that a Hilbert space X can be
decomposed into the orthogonal sum of a subspace Y and its complement Y K :
Corollary 1.5.7 (Orthogonal decomposition). Let X be a Hilbert space and
Y be a closed subspace. Then every vector x P X can be uniquely represented as
xy z, y P Y, z P Y K.
This orthogonal decomposition is usually abbreviated as
X Y ` Y K.
Definition 1.5.8 (Orthogonal projection). In the setting of Corollary 1.5.7,
the map
PY : X Ñ X, PY x y
is called the orthogonal projection in X onto Y .
Exercise 1.5.9. Show that the orthogonal projection PY is a linear
map. Check that ImpPY q Y and kerpPY q Y K . Also check that the
identity map IX on X can be decomposed as
IX PY PY K .
The proof of this result is based on its finite version, which may be called the
Pythagorean theorem in higher dimensions:
Lemma 1.6.7 (Pythagorean theorem). Let pxk q be an orthogonal system in a
Hilbert space X. Then for every n P N one has
¸n 2 ¸
n
(1.18)
xk }xk }2 .
k 1
k 1
Proof. Using orthogonality, we see that the left hand side of (1.18)
A¸
n ¸
n E ¸
n ¸
n
xk , xk xxk , xl y xxk , xk y
k 1
k 1
k,l 1
k 1
° Proof of Theorem 1.6.6. (i) ô (ii). By the Cauchy criterion, the series
x
k k converges if and only if its partial sums form a Cauchy sequence in X, i.e.
¸
m 2
(1.19)
xk Ñ0 as n, m Ñ 8.
k n
Note
°m that 2by Pythagorean theorem (Lemma 1.6.7), the quantity in (1.19) equals
kn }xk } . So using Cauchy°
criterion again we see that (1.19) is equivalent to the
convergence of the series k }xk }2 , as°
required.
(ii) ñ (iii). The scalar series k }xk }2 converges absolutely, therefore also
unconditionally (as we know from an analysis
° course). Hence, by the equivalence
of (i) and (ii) proved above, the series k xk converges unconditionally.
(iii) ñ (i) is trivial.
The last part of the theorem, identity (1.17), follows by taking limit in Pythagorean
identity (1.18). The theorem is proved.
°8
8 ak e
ikt
Exercise 1.6.8. For what coefficients ak does the series
converge in L2 rπ, π s?
k
1.6. FOURIER SERIES 31
Lec.11: 10/01
1.6.3. Fourier series.
Definition 1.6.9 (Fourier series). Consider an orthonormal system pxk q8
k1 in
a Hilbert space X and a vector x P X. The Fourier series of x with respect to pxk q
is the formal series ¸
xx, xk yxk .
k
The coefficients xx, xk y are called the Fourier coefficients of x.
In order to understand the convergence of Fourier series, we will first focus on
the finite case, and study the partial sums of Fourier series
¸
n
Sn pxq : xx, xk yxk .
k 1
This result along with the convergence criterion for orthogonal series, Theo-
rem 1.6.6, shows that Fourier series always converge.
Corollary 1.6.12. ° Let pxk q be an orthonormal system in a Hilbert space X.
Then the Fourier series k xx, xk yxk of every vector x P X converges in X.
12
°n Recall that the ( p
linear span of vectors x1 , . . . , xn is defined as Span x1 , . . . , xn q x
k1 ak xk : ak P C .
1.6. FOURIER SERIES 32
Proof. The first part follows from the Optimality Theorem 1.6.13, since by
completeness the orthogonal projection onto Spanpxk q X is the identity map in
X. Parseval’s identity follows from Fourier expansion (1.20), Pythagorean identity
(1.17) for orthogonal series, and the normalization condition }xk } 1.
Exercise 1.6.17. Prove that Parseval’s identity holds for an orthonor-
mal system pxk q if and only if pxk q is complete. Therefore the equality
cases of Bessel’s inequality hold exactly when the system is complete.
Now we describe some classical examples of complete sets and orthonormal
bases.
Example 1.6.18 (Monomials). Weierstrass approximation theorem states that
the system of monomials ptk q8 k0 is a complete system in C r0, 1s. We claim that
this is also a complete system in L2 r0, 1s.
Indeed, C r0, 1s is dense in L2 r0, 1s. This means that for every f P L2 r0, 1s
and ε ¡ 0, there exists g P C r0, 1s such that }f g }2 ¤ ε{2. By Weierstrass
approximation theorem, there exists h P Spanptk q8 k0 such that }g f }8 ¤ ε{2.
Hence }g f }2 ¤ }g f }8 ¤ ε{2, so by triangle inequality we conclude that
}f h}2 ¤ ε{2 ε{2 ε. We have proved that Spanptk q8k0 is dense in L2 r0, 1s as
required.
1.6. FOURIER SERIES 33
xk ptq ? eikt , k P Z,
1
(1.21)
2π
forms an orthonormal basis of L2 rπ, π s. Reformulating Theorem 1.6.16 in this
case, we obtain a basic result in classical Fourier analysis:
Theorem 1.6.20 (Classical Fourier series). Every function f P L2 rπ, πs can
be represented by its Fourier series
8̧ »π
f ptq fppk qeikt , fppk q f ptqeikt dt.
1
where
k 8 2π π
The coefficients fppk q are all finite; the Fourier series converges in L2 rπ, π s.
Example 1.6.21 (Trigonometric system). In a similar way we can show that
the trigonometric system considered in Example 1.6.4 is an orthonormal basis in
L2 rπ, π s (do this!) Therefore a version of Theorem 1.6.20 holds for the trigono-
metric system, and it reads as follows:
8̧
f ptq ak cospktq bk sinpktq
a0
2
k 1
where » »
1 π 1 π
ak f ptq cospktq dt, bk f ptq sinpktq dt.
π π π π
This again holds for every function f P L2 rπ, π s; the coefficients ak , bk are all
finite, and the Fourier series converges in L2 rπ, π s.
Lec.12: 10/04
Exercise 1.6.25. Prove that the spaces `8 and L8 are not separable
by constructing an uncountable separated subset A, i.e. such that
inf t}x y } : x, y P A, x yu ¡ 0.
For Banach spaces, separability is formally equivalent to a slightly stronger
property:
Lemma 1.6.26 (Separable spaces). A Banach space X is separable if and only
if it contains a system of vectors pxk q8
k1 whose linear span is dense in X, i.e.
Spanpxk q X.
Proof. Necessity. If X is separable, it contains a system of vectors pzk qnk1
whose linear span is dense in X. We construct pxk q inductively as a subset of pzk q.
Namely, we include z1 , and if zn 1 R pz1 , . . . , zn q we include zn for n 1, 2, . . .. By
construction, pxk q is linearly independent and Spanpx1 , . . . , xn q Spanpz1 , . . . , zn q
for all n 1, 2, . . . Letting n Ñ 8 we conclude that Spanpx1 , . . . , xn q Spanpz1 , . . . , zn q
X as required.
Sufficiency.
°n If Spanpxk q is dense in X, so is the set of all finite linear combina-
tions k1 ak xk with ak P Q, which is a countable set. The lemma is proved.
°
Note that every x P X has the form x k ak xk for some (Fourier) coefficients ak ,
so the definition makes sense. Also, by Parseval’s identity,
¸ 2 ¸ ¸ 2
(1.23)
a k xk |ak |2
ak yk .
k k k
Remark 1.6.31. It follows from Theorem 1.6.30 that }T x} }x} for all x P X.
Thus by linearity,
}T x T y} }x y} for all x, y P X.
Thus T preserves all pairwise distances; hence the name “isometry”.
Remark 1.6.32. Since `2 and L2 r0, 1s are separable Hilbert spaces, it follows
that every Hilbert space is isometric to `2 and L2 r0, 1s.
Now consider the weight wptq ?12π et {2 , i.e. the standard normal
2
dt k
On the other hand, n}xxnn } Ñ 0 as the norm of these vectors equals 1{n. This implies
that f is not continuous.
(iii) For every hyperplane H E there exists a linear functional f 0 such that
ker f H.
Proof. (i) Follows from a linear version of the fundamental theorem on ho-
momorphisms, Exercise 1.1.24. Indeed, the injectivization f˜ : E { ker f Ñ C of f
establishes a linear bijection (isomorphism) between E { ker f and the range C of f .
Thus dimpE { ker f q dimpCq 1, so ker f is a hyperplane in E.
(ii) Since ker f ker g : H, the injectivizations f˜, g̃ : E {H Ñ C are linear
functionals on the one-dimensional space E {H. A moment’s thought yields that
such linear functionals must be equal up to some constant factor a, i.e. f˜
ag̃. Then f˜rxs ag̃ rxs for all x P E. On the other hand, by construction of
injectivization, f pxq f˜rxs and g pxq g̃ rxs. Therefore f pxq ag pxq as required.
(iii) Since dimpE {H q 1, we have
E {H tarx0 s : a P Cu
for some x0 P E. Let x P E be arbitrary; then rxs arx0 s for some a apxq P C,
which implies x ax0 h for some h P H. Let us define f on E by f pxq a. Then
f is a linear functional (check!), and clearly ker f H.
Remark 2.1.15. Using injectivization of f , one can show that the converse also
holds. So, a linear functional f is bounded if and only if ker f is closed. It follows
that the kernel of a linear functional is either closed or dense in X. (Why?)
Lec.14: 10/08
is a bounded linear functional on the space L2 pµq, and therefore also on the space
L2 pµ ν q. By Riesz representation theorem, there exists h P L2 pµ ν q such that
» » » »
(2.4) f dµ f h d pµ νq f h dµ f h dν for all f P L2 pµ ν q.
We claim that
By Monotone Convergence Theorem, one can show that (2.5) holds for arbitrary
pµ ν q-measurable functions f such that f ¥ 0 pµ ν q-a.e. (Indeed, consider the
truncation fn ptq : minpf ptq, nq and let n Ñ 8.) The convention is that if one side
of (2.5) is infinite then the other is infinite, too.
Now, given a measurable set A, we choose f so that f h 1A . In other words,
we consider
f :
1A
h
and apply the identity in (2.5). We obtain
»
1h
ν pAq dµ.
A h
Lec.15: 10/11
2.2.3. The dual of Lp . A version of the representation theorem for L2 , Corol-
lary 2.2.3, holds in fact for all Lp spaces. In short, it states that Lp Lp1 where p
and p1 are conjugate exponents as in Hölder’s inequality, i.e.
1
p p1
1
1, 1 p, p1 8,
and p1 8 if p 1. The rigorous statement is the following:
Theorem 2.2.5 (Lp Lp1 ). Consider the space Lp Lp pΩ, Σ, µq with finite
of σ-finite measure µ, and where 1 ¤ p 8. Let p1 be the conjugate exponent of p.
(i) For every weight function g P Lp1 , integration with weight
»
Gpf q : f g dµ, f P Lp
is a bounded linear functional on Lp , and its norm is }G} }g }p1 .
(ii) Conversely, every bounded linear functional G P Lp can be represented as
integration with weight for some unique weight function g P Lp1 . Moreover,
}G} }g}p1 .
We shall prove Theorem 2.2.5 for the particular case of spaces `p . The general
case is somewhat similar, and it relies on an application of Radon-Nykodim theorem.
For the spaces `p , Theorem 2.2.5 reads as follows:
Corollary 2.2.6 (`p `p1 ). Let 1 ¤ p 8 and let p1 be the conjugate
exponent of p.
(i) For every y P `p1 , summation with weight
8̧
Gpxq : xk yk , x P `p
k 1
is a bounded linear functional on `p , and its norm is }G} }y }p1 .
(ii) Conversely, every bounded linear functional G P `p can be represented as
summation with weight for some unique weight y P `p1 . Moreover, }G} }y }p1 .
Proof. We will only prove the case where 1 p, p1 8; the case p 1,
p1 8 is an exercise.
(i) By Hölder’s inequality, we have
¸
|Gpxq| xk yk ¤ }x}p }y }p1 .
k
It follows that F is a bounded linear functional on `p , and }G} ¤ }y }p1 . To prove
the converse inequality, note that Hölder’s inequality is sharp. Namely, for every
y P `p1 there exists x P `p such that
¸
|Gpxq| xk yk }x}p }y }p1 .
k
Indeed, one can check that this holds for x pxk q defined as
1
xk ei Argpyk q |yk |p 1 .
For this x, it follows that }G} ¥ }y }p1 , so part (i) of the theorem is proved.
1The result can be extended (by decomposition) to σ-finite measures.
2.2. REPRESENTATION THEOREMS FOR LINEAR FUNCTIONALS 46
2.2.4. The dual of C pK q. Finally, we state without proof the following char-
acterization of bounded linear functionals on C pK q.
Theorem 2.2.8 (C pK q ). Consider the space C pK q where K is a compact
topological space.
(i) For every Borel regular signed measure2 µ on K, integration
»
Gpf q : f dµ, f P C pK q
is a bounded linear functional on C pK q, and its norm is the total variation
}G} |µ|pK q.
(ii) Conversely, every bounded linear functional G P C pK q can be represented
as integration with respect to a unique Borel regular signed measure µ on K.
Moreover, }G} |µ|pK q.
Lec.17: 10/15
2.3.3. Supporting functionals. Hahn-Banach theorem has a variety of con-
sequences, both analytic and geometric. One of the basic tools guaranteed by
Hahn-Banach theorem is the existence of a supporting functional f P X for every
vector x P X.
Proposition 2.3.7 (Supporting functional). Let X be a normed space. For
every x P X there exists f P X such that
}f } 1, f pxq }x}.
The functional f is called the supporting functional of x.
Proof. Consider the one-dimensional subspace X0 Spanpxq, and define a
functional f0 P X0 by
f0 ptxq t}x}, t P R or C.
Then }f0 } 1. An extension f P X of f0 guaranteed by Hahn-Banach theorem
clearly satisfies the conclusion.
Generally it is not true that every functional f attains its norm on some vector x,
i.e. that the supremum above can be replaced by the maximum.
Exercise 2.3.9. Construct a bounded linear functional on C r0, 1s which
does not attain its norm.
However, every vector x does attain its norm on some functional f P X ,
namely the supporting functional. This immediately follows from Proposition 2.3.7:
Corollary 2.3.10. For every vector x in a normed space X, one has
2.3.4. Second dual space. Let X be a normed space as usual. The func-
tionals f are designed to act on vectors x P X via f : x ÞÑ f pxq. Vice versa, we
can say that vectors x P X act on functionals f P X via
x P X
}x}X }x}.
We demonstrated that there exists a canonical embedding of X into into X .
We summarize this as follows.
Theorem 2.3.12 (Second dual space). Let X be a normed space. Then X can
be considered as a linear subspace of X . For this, a vector x P X is considered
as a bounded linear functional on X via the action
x:f ÞÑ f pxq, f P X .
Example 2.3.13. As we know from Section 2.2.3, c0 `1 while `1 `8 , so
c
0 `8 .
The space c0 of sequences converging to zero is indeed canonically embedded into
the larger space `8 of all bounded sequences (and with the same sup-norm).
Proof. Let
r distpA, B q : aPA,infbPB }a b}.
By the assumptions, r ¡ 0. (Why?) Therefore, the open r{3-neighborhoods Ar{3
of A and Br{3 of B are disjoint, open and convex sets. Applying Theorem 2.3.25,
we obtain a functional f P X that separates the neighborhoods:
sup f paq ¤ inf f pbq.
P
a A r {3 bPBr{3
Remark 2.3.29. Suppose that in Theorem 2.3.24, the set K is either open (i.e.
as stated) or closed. Then the strict separation holds:
f pxq f px0 q, x P K.
Indeed, for open sets this follows from Theorem 2.3.25, while for closed sets this
follows from Corollary 2.3.27.
2.3.7. Convex sets are intersections of half-spaces.
Corollary 2.3.30. Every closed convex subset K of a normed space X is the
intersection of all (closed) half-spaces that contain K.
Recall that the half-space is what lies on one side of a hyperplane; therefore
half-spaces have the form
tx P X : f pxq ¤ au
for some f P X , a P R. See the picture illustrating Corollary 2.3.30.
Exercise 2.3.32. [Convex sets that can not be separated] Show that
the openness assumption in Theorem 2.3.25 is essential.
To this end, consider the linear space P of all polynomials in one
variable and with real coefficients. Let the subset A consist of polyno-
mials with negative leading coefficient, and let the subset B consists of
polynomials with all non-negative coefficients. Show that A and B are
disjoint convex subsets of P, and that there does not exist a nonzero
linear functional f on P such that
f paq ¤ f pbq for all a P A, b P B.
(Hint: assume that for some C P R one has f paq ¤ C ¤ f pbq, a P A, b P B;
deduce from 0 P B that C ¤ 0 and by considering monomials that C ¥ 0.
2.4.2. Space of operators. Let X and Y be normed spaces. The space of all
bounded linear operators T : X Ñ Y equipped with the operator norm is denoted
LpX, Y q.
As an example, the dual space is a space of operators that map to scalars, i.e.
X LpX, Rq.
Proposition 2.4.6. LpX, Y q is a normed space. Moreover, if Y is a Banach
space then LpX, Y q is also a Banach space. In particular, the dual space X is
always a Banach space, even if X is incomplete.
Proof. The norm axioms are straightforward. (Check!) To prove complete-
ness, let Tn P LpX, Y q be a Cauchy sequence, i.e.
}Tn Tm } Ñ 0, n, m Ñ 8.
Applying this to an arbitrary x P X and noting that
}Tn x Tn x} ¤ }Tn Tm } }x} Ñ 0, n, m Ñ 8
we see that pTn xq is a Cauchy sequence in Y . By the completeness of Y it converges.
Define the map T as
T x : lim Tn x.
n
We claim that T is the limit of Tn in LpX, Y q, which would complete the proof.
It is easy to check that T : X Ñ Y is a linear operator. (Check!)
To show that T is bounded, we choose an arbitrary x P X and use continuity
of the norm:
}T x} lim }Tn x} ¤ sup }Tn }}x}.
n n
Since a Cauchy sequence is always bounded (why?), supn }Tn } 8. It follows that
T is a bounded linear operator, i.e. T P LpX, Y q.
It remains to show that Tn Ñ T in LpX, Y q, i.e. in the operator norm. Since
T is Cauchy, for every ε ¡ 0 there exists a N P N such that
}Tn Tm } ε for n, m ¡ N.
Applying this to an arbitrary x P X we obtain
}Tn x Tm x} ε}x} for n, m ¡ N.
2.4. BOUNDED LINEAR OPERATORS 58
Letting m Ñ 8, we get
}Tn x T x} ε}x} for n ¡ N.
Since x is arbitrary it follows that
}Tn T } ε for n ¡ N.
This means that Tn Ñ T in LpX, Y q as required.
We can view this definition as a continuous version of (2.10), where kernel k pt, sq
can be considered as a continuous version of matrix. The operator T defined this
way is called Hilbert-Schmidt integral operator with kernel k pt, sq.
Proposition 2.4.8. A Hilbert-Schmidt integral operator T : L2 r0, 1s Ñ L2 r0, 1s
with kernel k pt, sq P L2 pr0, 1s2 q is bounded. Specifically,
}T } ¤ }k}2 .
2.4. BOUNDED LINEAR OPERATORS 59
Differentiating yields
Dek pikqek , k P Z.
Hence
}Dek }2 k}ek }2 k, k P Z.
Letting k Ñ 8 implies that D is an unbounded operator.
This simple example suggests that various differential operators are unbounded
even on good function spaces. They may be studied through their inverses, which
are bounded integral operators.
(2.14) Dn pθq
¸
n
eikθ
pn 2 qθ
sinsin
1
1
k n θ 2
As we know from Section 2.3, every bounded linear functional can be extended
from either dense or closed subspace to the whole space. For dense subspaces, we
2.4. BOUNDED LINEAR OPERATORS 62
can extend by continuity, while for closed subspaces the extension is guaranteed by
Hahn-Banach theorem.
For general linear operators, extension by continuity holds with the same proof
as in Proposition 2.3.2:
Proposition 2.4.20 (Extension by continuity). Let X0 be a dense subspace of
a normed space X, and Y be a Banach space. Then every operator T0 P LpX0 , Y q
admits a unique extension T P LpX, Y q. Moreover, }T } }T0 }.
Unfortunately, extension from a closed subspace is not always possible, and
Hahn-Banach theorem does not generalize to bounded linear operators. There is
a simple geometric description of the situations when such extensions are possible.
To state it we need a general notion of projections in normed space (not necessarily
orthogonal).
Definition 2.4.21 (Projection). Let X0 be a closed subspace of a normed
space X. An operator P P LpX, X q is called a projection in X onto X0 if
(i) P pX q X0 ;
(ii) P x x for all x P X0 , i.e. P |X0 IX0 .
Example 2.4.22. Any orthogonal projection in a Hilbert space is clearly a
projection in this sense. However, even in a Hilbert space there is a plenty of
non-orthogonal projections. (Construct one in a two-dimensional space.)
The following observation characterizes the subspaces from which extensions of
linear operators are possible.
Proposition 2.4.23 (Extensions of operators and projections). Let X0 be a
closed subspace of a normed space X. Then the following are equivalent:
(i) There exists a projection in X onto X0 . In this case we say that X0 is a
complemented subspace of X;
(ii) For every normed space Y , every operator T0 P LpX0 , Y q admits an extension
T P LpX, Y q.
Proof. Assume P is a projection in X onto X0 . Then for every operator
T0 P LpX0 , Y q, the operator T : T0 P P LpX0 , Y q is an extension.
Vice versa, consider the identity map I : X0 Ñ X0 . Its extension P : X Ñ X0
is clearly a projection in X onto X0 .
Since every closed subspace in a Hilbert space is complemented, the extension
problem in Hilbert spaces always has a positive solution.
Unfortunately, in general normed spaces there may be uncomplemented sub-
spaces. Here is a synopsis of some of the known results, without proof:
Theorem 2.4.24 (Complemented subspaces). In what follows, X0 stands for
a closed subspace of a Banach space X.
(i) If either dim X0 8 or codim X 8 then X0 is complemented in X.
(ii) c0 is not complemented in `8 [10].
(iii) Every complemented subspace X0 of X is isomorphic to X if X is one of the
spaces `p , p P r1, 8s or c0 . These results are due to Pelczynski and Linden-
strauss [9, 6].
(iv) Every Banach space that is not isomorphic to a Hilbert space has an uncom-
plemented subspace. This is a theorem of Lindenstrauss and Tzafriri [7].
2.4. BOUNDED LINEAR OPERATORS 63
}T }
as required.
2.4. BOUNDED LINEAR OPERATORS 64
Exercise 2.4.29. Let R and L denote the right and left shift operators
on `2 . Prove that R L.
Im T ` ker T .
H
ker T pIm T qK .
Deduce that
pker T qK
Im T .
Give an example of a linear operator such that pker T qK Im T .
2.4. BOUNDED LINEAR OPERATORS 65
Proof. It suffices to prove the result for x P ker P and for x P Im P , because
then the result will follow for all x P H by the orthogonal decomposition H
ker P ` Im P . (Check!)
For x P Im P the result is trivial because in this case U n x U x x and
P x x. So let x P ker P . We will first find a convenient representation of ker P .
By definition, Im P kerpI U q kerpI U q because for unitary operators,
U x x if and only if U x x. (Check!) Therefore, using the duality between
kernels and images, Corollary 2.4.33, we have
ker P pkerpI U qqK ImpI U q.
Therefore, every x P ker P can be approximated arbitrarily well by vectors of the
form pI U qy. For x pI U qy, we arrive at a telescoping sum
N 1
1 ¸ n
U x p x U N xq Ñ 0 Ñ 8.
1
as N
N n0 N
(2.17)
|tn ¤ N : mod 2π P Au|
nθ pAq
Ñ µ2π as N Ñ 8.
N
where µ denotes the Lebesgue measure.
2.4. BOUNDED LINEAR OPERATORS 66
Theorem 2.4.35 implies a very general form of Weyl’s ergodic theorem, for an
arbitrary ergodic measure-preserving transformation of a probability space (instead
of an irrational rotation of the circle).
Definition 2.4.37 (Ergodic transformation). Let pΩ, Σ, µq be a probability
space. A transformation T : Ω Ñ Ω is called measure-preserving if
µpT 1 pAqq µpAq
for all measurable subsets A Ω.4 A one-to-one, measure preserving transforma-
tion T is ergodic if the only functions f P L2 pΩ, Σ, µq which satisfy f pT ω q f pω q
for almost all ω P Ω are the constant functions.
Exercise 2.4.38. Show that T is ergodic if and only if for all measur-
able subset A Ω, T 1 pAq A implies µpAq 0 or µpAq 1.
Weyl’s ergodic theorem (2.17) is a partial case of Theorem 2.4.39 for Ω being
the unit circle, T being the rotation of the circle by an irrational angle θ (why is T
ergodic?) and for f 1A .
p q tω P Ω :
4Here T 1 A Tω P Au is the preimage of A under T .
2.4. BOUNDED LINEAR OPERATORS 67
Show that the intergal operator (2.11) T : L2 r0, 1s Ñ L2 r0, 1s with kernel
k pt, sq is bounded, and5 a
}T } ¤ M1 M2 .
Exercise 2.4.42. [Multiplication operator] Consider a multiplier func-
tion k ptq P C r0, 1s, and define a linear operator T : C r0, 1s Ñ C r0, 1s by
pT f qptq kptq f ptq.
Show that T is a bounded linear operator and compute its norm.
69
3.1. OPEN MAPPING THEOREM 70
B K1
2
B,
for the right side is the ε-neighborhood of K in Y . Iterating this inclusion gives
1 1
B K 2
K
2
B K
1
2
K
1
4
B
K 1
2
K
1
4
K
1
2
B K
1
2
K
1
4
K
1
8
B
Therefore1
B K 1
2
K
1
4
K
1
8
K
By perfect convexity (check!), we have
B K K.
1 1 1 1 1
K K K
2 2 4 8 16
This proves the lemma.
Exercise 3.1.5. Verify the steps of the above proof where we used
Minkowski sums and series of sets.
Now we are ready to complete the proof of the open mapping theorem. By
Lemma 3.1.4, it suffices to show that K T BX is perfectly convex. This is easy to
°
° All sums and series of sets are in Minkowski sense. The sum of sets is defined as k Ak
1
t k ak : ak P Ak u. The same for infinite sums (series), where we insist on the convergence of
°
k ak .
3.1. OPEN MAPPING THEOREM 71
In other words, the norms } } and ~ ~ are equivalent if the identity operator
(3.4) I : pE, } }q Ñ pE, ~ ~q
is an isomorphism.
Exercise 3.1.9. Show that two norms on a linear vector space E are
equivalent if and only if they generate the same topology on E.
It follows that for every two n-dimensional normed spaces X and Y , one has
dpX, Y q ¤ n.
(Why?) E. Gluskin [4] proved in 1981 that this upper bound is asymptotically
sharp, i.e. there exists an absolute constant c ¡ 0 such that, for every n P N one
can construct n-dimensional normed spaces Xn and Yn with
dpXn , Yn q ¥ cn.
Corollary 3.1.17. (i) Every finite dimensional normed space is a Banach
space.
(ii) Every finite dimensional subspace of a normed space is closed.
(iii) Every linear operator on a finite dimensional normed space is bounded.
(iv) Every two norms on a finite dimensional normed space are equivalent.
Proof. (i) Since an n-dimensional normed space X is isomorphic to the com-
plete space `n2 , X itself is complete.
(ii) Let Y be a finite-dimensional subspace of a normed space X. Then Y
is complete by part (i). Therefore Y is closed. (Indeed, if a sequence pyn q Y
converges to x P X then pyn q is Cauchy in Y , hence its limit x is in Y .)
(iii) By Proposition 2.4.7, every linear operator on `n2 is bounded, }T } ¤ }T }HS .
Since an n-dimensional normed space X is isomorphic `n2 , the same is true for linear
operators on X.
(iv) Let } } and ~ ~ be two norms on X. By part (iii), the identity map
I : pX, } }q Ñ pX, ~ ~q as well as its inverse are bounded. This completes the
proof.
Exercise 3.1.18. [Operators from finite dimensional spaces] Let X and
Y be normed spaces, and X be finite dimensional. Show that every
linear operator T : X Ñ Y is bounded. (Hint: identify X with `n2 by some
isomorphism, and show boundedness of T similarly to the argument of
Theorem 3.1.13.)
Lec. 24: 11/3
where C 1 r0, 1s is considered as a subspace of C r0, 1s, i.e. with respect to sup-norm.
Lemma 3.2.4. The differential operator T has closed graph.
Proof. Let fn Ñ f and T fn fn1 Ñ g in C r0, 1s, i.e. uniformly. Then, by
the theorem on differentiation under the limit,3 plimn fn q1 limn fn1 , i.e. g T f .
This completes the proof.
Nevertheless, as we know the differential operator is unbounded. (Why?) This
does not contradict the closed graph theorem, because C 1 r0, 1s is not complete
under the sup-norm. If we consider C 1 r0, 1s under its natural norm }f }8 }f 1 }8
in which it is a Banach space (Exercise 1.3.12), then the differential operator will
obviously be bounded.
p
3The theorem on differentiation under the limit states that lim f q1 limn fn1 provided
that fn1 converges uniformly and fn t0 converges for some point t0 .
n n
p q
3.3. PRINCIPLE OF UNIFORM BOUNDEDNESS 76
T i dtd
with domain
Dom T tf P L2 r0, 1s : f P C 1 r0, 1s, f p0q f p1q 1u.
3.3. Principle of uniform boundedness
Lec.25: 11/5
3.3.1. Statement and proof. The principle of uniform boundedness is a
result due to Banach and Steinhaus.
Theorem 3.3.1 (Principle of uniform boundedness). Let X and Y be Banach
spaces. Consider a family of bounded linear operators T LpX, Y q. Assume that
T is pointwise bounded, i.e.
(3.8) sup }T x} 8 for every x P X.
P
T T
Then T is uniformly bounded, i.e.
sup }T } 8.
P
T T
Note that the reverse direction is trivially true – uniform boundedness implies
pointwise boundedness. (Why?)
3.3. PRINCIPLE OF UNIFORM BOUNDEDNESS 77
Baire category theorem implies that one of Xn is not a nowhere dense subset of
X. Since Xn is closed (why?), Xn has nonempty interior. Summarizing, we have
shown that there exist n P N, x0 P X and ε ¡ 0 such that
x0 εBX .
Xn
By symmetry of Xn , we also have Xn
x0 εBX . Hence by convexity of Xn ,
we have
Xn
εBX .
(Check!) By definition of Xn this means that for x P X,
}x} ¤ ε implies sup }T x} ¤ n.
T T P
It follows that for every x P X,
sup }T x} ¤ }x}.
n
P
T T ε
This implies that
sup }T } ¤ 8
n
P
T T ε
as required.
Exercise 3.3.2. Check that the sub-level sets Xn in the proof above
are closed, convex and symmetric. (All these properties were used in
the argument).
Here again the reverse statement is trivially true – (strong) boundedness triv-
ially implies weak boundedness.
Proof. We embed X into X using the canonical embedding that we studied
in Theorem 2.3.12. So we consider vectors x P A as bounded linear functionals on
X acting as xpf q : f pxq, f P X . Rewriting the weak boundedness assumption
as supxPA |xpf q| 8 for f P X , we may understand this assumption as point-
wise boundedness of the family A X LpX , Rq. The principle of uniform
boundedness implies that supxPA }x}X supxPA }x}X 8, as required.
Remark 3.3.6. Using Corollary 3.3.5, one can weaken the assumption (3.8) in
the principle of uniform boundedness to the following one:
sup |f pT xq| 8 for every x P X, f P Y .
P
T T
(Why?)
3.3.3. Application to convergence of Fourier series. A basic and classi-
cal question in Fourier analysis is – when does Fourier series of a function f on an
interval converge to f ?
Hilbert space technique provides a complete answer to this question in the
space L2 . As we know from Theorem 1.6.20, the Fourier series of every function in
L2 rπ, π s converges to f in the L2 -norm.
In function spaces other than L2 , the answer to this problem is often nontrivial
and even negative. Unfortunately, such is the situation in the space of continuous
functions C rπ, π s. There exist continuous functions f whose Fourier series do
not converge in C rπ, π s (i.e. uniformly). This follows from a somewhat stronger
result, which in turn is a consequence of the principle of uniform boundedness:
Theorem 3.3.7 (Divergent Fourier series). There exists a function f P C rπ, πs
whose partial Fourier sums
¸
n
pSn f qptq fppk qeikt
k n
forms an unbounded sequence of complex numbers at t 0. In particular, the
Fourier series of f is unbounded (hence divergent) at t 0.
Proof. Recall from (2.13) that the partial Fourier sums can be represented
via convolution with Dirichlet kernel Dn :
»π
sinpn 21 qθ
pSn f qptq 2π1 Dn pt sqf psq ds, where Dn pθq .
π sin 21 θ
We are interested in the behavior of pSn f qp0q. These are obviously linear functionals
on C rπ, π s, which we denote
»π
φn pf q : pSn f qp0q Dn psqf psq ds.
1
2π π
(We used that Dn is an even function).
3.3. PRINCIPLE OF UNIFORM BOUNDEDNESS 79
Remark 3.3.9 (Convergence of Fourier series). (i) Analyzing the proof of The-
orem 3.3.7 one can show that continuous functions whose Fourier series con-
verge pointwise are rare. Precisely, the set of such functions is of first Baire
category in C rπ, π s (i.e. it is a countable union of nowhere dense sets).
(ii) Nevertheless, for every continuous function f , and even for f P Lp rπ, π s,
p ¡ 1, Fourier series converges to f almost everywhere. So the set of points
of divergence is always small. This is a deep result of L. Carleson [2].
(iii) For functions in L1 , Carleson result is generally false. Kolmogorov (apparently
when he was 19-21!) constructed a function f P L1 rπ, π s whose Fourier series
diverges everywhere.
(iv) If f is differentiable at a point t, then Fourier series of f converges to f at t.
This is called Dirichlet-Dini condition.
3.3. PRINCIPLE OF UNIFORM BOUNDEDNESS 80
However, the coefficients ak ak pεq may depend on ε. The limit limεÑ0 ak pεq
generally does not exist, as will be clear from the next exercise. In contrast, the
basis property guarantees that one can achieve approximation (3.11) with the same
coefficients ak independent of ε by only increasing the number of terms n in the
linear combination.
Exercise 3.3.12. [Completeness and basis property] In a Hilbert space
H, find a complete and linearly independent sequence pxk q which is not a
Schauder basis. Hint: construct xk so that it converges to some nonzero
vector in H; show that such sequences are never Schauder bases.
Example 3.3.15 (Basis of the space of continuous functions). In C r0, 1s, the
natural candidates fail to be Schauder bases. The Fourier basis is not a Schauder
basis – otherwise this would imply that the Fourier series of every continuous func-
tion would converge in C r0, 1s (why?), which would contradict Theorem 3.3.7.
The sequence of monomials 1, t, t2 , . . . is not a Schauder basis of C r0, 1s either.
(Prove this!)
The most known Schauder basis of C r0, 1s is the so-called Schauder system of
wavelets. Its mother wavelet φptq is obtained by integration of Haar mother wavelet,
i.e. #
»t
t P r0, 1{2q
φptq hpsq ds
t,
0 1 t, t P r1{2, 1q
Then we consider the translates and dilates of the mother wavelet:
φkl ptq φp2k t lq, k 0, 1, 2, . . . , l 0, 1, 2, . . . , 2k 1.
Together with the constant function 1, the system of functions φkl ptq is called the
Schauder system, see the picture. It forms a Schauder basis on C r0, 1s. (Check!)
The next useful result states that pointwise convergence of operators implies
uniform convergence on compacta. We say that a sequence of operators Tn P
LpX, Y q between normed spaces X and Y converges pointwise to some T P LpX, Y q
if
Tn x Ñ T x for all x P X.
3.4. COMPACT SETS IN BANACH SPACES 85
Even though weak convergence is generally strictly weaker than strong con-
vergence, there are several useful ties between weak and strong properties. Weak
convergence clearly implies weak boundedness, which in turn implies strong bound-
edness by a consequence to the principle of uniform boundedness (Corollary 3.3.5):
Proposition 3.5.3. Weakly convergent sequences in Banach spaces are bounded.
Moreover, we have a good control of the weak limit, given in the next two
results.
Proposition 3.5.4. If xn ÝÑ
w
x in a normed space then }x} ¤ lim inf n }xn }.
Proof. Let f P X be a supporting functional of x, i.e. }f } 1, f pxq }x}.
Then f pxn q ¤ }xn } for all n. Taking lim inf of both sides, we conclude that
lim inf }xn } ¥ lim inf f pxn q f pxq }x}
n n
as required.
p q
6Recall that conv A is the smallest convex set containing A, see Exercise 1.2.24.
3.5. WEAK TOPOLOGY 88
and
f pxk q Ñ f pxq for every f P A.
Proof. Nesessity follows by Proposition 3.5.3.
Sufficiency will be proved by a standard approximation argument. Consider
arbitrary g P X and ε ¡ 0. By density, we can choose f P A such that }g f } ¤ ε.
Then
lim sup |g pxk xq| ¤ lim sup |f pxk xq| lim sup |pg f qpxk xq|
k k n
¤ 0 }g f } lim supp}xk } }x}q ¤ M ε
k
where M supk }xk } }x} 8 by the boundedness assumption. Since ε ¡ 0
is arbitrary, we conclude that lim supk |g pxk xq| 0. Hence g pxk q Ñ g pxq as
required.
Theorem 3.5.7 (Weak convergence in c0 and `p ). Let X c0 or X `p ,
p P p1, 8q. Then xk Ý Ñ x in X if and only if the sequence pxk q is bounded and
w
for every Borel regular signed measure µ. On the other hand, our assumptions are
that the sequence of functions xn ptq is uniformly bounded and it converges to xptq
pointwise. The Lebesgue dominated convergence theorem implies (3.16).
Proof. One notices that the set of characteristic functions 1ra,bs ptq for ra, bs
r0, 1s spans the set of step functions, which is dense in pLp q Lp1 .(Why?) The
argument is finished similarly to Theorem 3.5.7.
Remark 3.5.18. The cylinders are clearly open in the strong topology. Hence
the weak topology is weaker than the strong topology.
Remark 3.5.19. In infinite-dimensional spaces X, the cylinders are rather large
as they contain subspaces of finite codimension tx P X : fk px x0 q 0, k
1, . . . , N u. This shows that in infinite dimensions, weakly open sets are unbounded.
Exercise 3.5.20. Prove that in an infinite dimensional normed space
X, weak topology is strictly weaker than the strong topology. Why
does not this contradict Schur property of X `1 mentioned in Re-
mark 3.5.14?
Nevertheless, some weak and strong properties are equivalent. For example,
weak boundedness and strong boundedness are equivalent. This follows from the
principle of uniform boundedness, see Corollary 3.3.5. (Check that what we called
“weak boundedness” is indeed boundedness in the weak topology).
Also, weak closedness and strong closedness are equivalent for convex sets:
Proposition 3.5.21 (Weak closedness). Let K be a convex set in a normed
space X. Then K is weakly closed if and only if K is (strongly) closed.
Proof. Necessity is trivial.
Sufficiency. Assume K is closed and convex. By Corollary intersection of half-
spaces to Hahn-Banach theorem, K is the intersection of the closed half-spaces that
contain K. Each closed half-space has the form
Af,a tx P X : f pxq ¤ au
for some f P X and a P R. Hence Af,a is weakly closed.8 The intersection K of
the closed half-spaces is therefore automatically weakly closed.
7That is, sup x
k } }8 8
k .
8Indeed, A
f,a is the preimage of the closed subset p8, as under the function f : X Ñ R,
which is continuous by the definition of weak topology.
3.6. WEAK TOPOLOGY. BANACH-ALAUGLU’S THEOREM 91
Example 3.6.3 (Dirac delta function). Recall that we understand Dirac delta
function δ ptq as the point evaluation functional at zero, see Example 2.1.3. Equiv-
alently, Dirac delta function may be identified with the probability measure on R
with the only atom at the orgigin. Therefore Dirac delta function is the weak limit
of uniform measures on r n1 , n1 s as n Ñ 8.
This gives a natural way to approximate Dirac delta function δ ptq (which does
not exists as a function on R) by genuine functions δn ptq, which are the probability
distribution functions of the uniform measures on r n1 , n1 s, see the picture.
3.6. WEAK TOPOLOGY. BANACH-ALAUGLU’S THEOREM 92
Each set Bx,y,a,b is the preimage of the weak closed set t0u under the map f ÞÑ
f pax by q af pxq bf py q which, as we know, is continuous in the product topology.9
Therefore all sets Bx,y,a,b are weak closed, and so is their intersection BX . This
completes the proof.
3.6.3. Universality of space C pK q. As an application of Banach-Alaoglu’s
Lec. 30: 11/17
9Recall that the point evaluation maps are continuous in the product topology.
10A little disclaimer is that the compact topological space K may depend on X; otherwise
the result is false for spaces X of too large cardinality.
CHAPTER 4
(We can do this by continuity of the kernel k pt, sq.) Now, for every f P BC r0,1s , we
obtain by triangle inequality that
»1
p qp q p qp q ¤
T f t1 T f t2 |kpt1 , sq f pt2 , sq| |f psq| ds ¤ ε
0
as |f psq| ¤ 1 for all s. This shows that the set K is equicontinuous, and therefore
precompact.
Exercise 4.1.5. Show that Volterra operator (2.12) is compact on
C r0, 1s, even though its kernel is discontinuous. See Exercise 4.1.19 for a
more general result.
For what multiplier sequences pλk q is the operator T (a) well defined?
(b) bounded? (c) compact?
The quantity
8̧
1{2
}T }HS : }T xk } 2
k 1
is called the Hilbert-Schmidt norm of T .
Exercise 4.1.12. For operators in finite dimensional spaces T : Cm Ñ
C , (equivalently, for m n matrices), we already gave a definition of
n
Remark 4.1.18 (For future). Consider proving the reverse direction in Schauder’s
theorem. Also consider proving that compact operators map weak Cauchy se-
quences to strongly convergent.
λx T x b
T
d
dt
acting, for example, on C 1 pCq. To compute the spectrum of T , we solve the ordinary
differential equation u1 λu. The solution has the form
uptq Ceλt .
2Recall however that the eigenvalues do not need to form a basis of Cn . The dimension of
less
the span of eigenvectors corresponding to a given eigenvalue (the eigenspace) may be strictly
than the multiplicity of that root. This happens, for example, for the Jordan block T
λ 1
0 λ
.
An orthonormal basis of eigenvectors exists if and only if T is normal, i.e. T T
TT .
3In the future, we will often say “invertible” instead of ‘invertible as a bounded linear
operator”.
4.3. SPECTRUM OF A BOUNDED LINEAR OPERATOR 102
Remark 4.3.7. If Dirac delta function δ ptq was a genuine function in L2 , then
its translates δλ ptq : δ pt λq would be the eigenvectors of the multiplication
operator on L2 :
T δλ λδλ
and δλ would be the eigenfunctions of T . The situation would be similar to the
discrete multiplication operator from Example 4.3.5.
Example 4.3.8 (Shift operator). Consider the right and left shift operators on
`2 , acting on a vector x px1 , x2 , . . .q as
Rpxq p0, x1 , x2 , . . .q, Lpxq px2 , x3 , . . .q.
Since R is clearly injective but Im R is not dense in `2 (why?), 0 is in the residual
spectrum of R. One can show that
σp pRq H, σc pRq tλ P C : |λ| 1u, σr pRq tλ P C : |λ| 1u;
σp pLq tλ P C : |λ| 1u, σc pRq tλ P C : |λ| 1u, σr pRq H.
Exercise 4.3.9. Prove the claims about the spectra of shift operators
made in Example 4.3.8.
8̧ 8̧
pI S q Sk S k pI Sq I
k 0
k 0
as telescoping series. (Check!) Finally,
8̧
}pI S q1 } ¤ }S }k ¤ 1 1}S } .
k 0
This completes the proof.
4.4. PROPERTIES OF SPECTRUM. SPECTRUM OF COMPACT OPERATORS. 104
Remark 4.4.8. It follows that for every functional f P LpX, X q , the function
f pRpµqq is a usual (i.e. complex-valued) analytic function on ρpT q.
Theorem 4.4.9. The spectrum σ pT q is a nonempty set.
4.4. PROPERTIES OF SPECTRUM. SPECTRUM OF COMPACT OPERATORS. 105
Proof. We shall deduce this result from Liouville’s theorem in complex anal-
ysis.4 To this end, assume that σ pT q H, hence ρpT q C and the resolvent Rpλq
is an entire function (i.e. analytic on the whole complex plane).
Claim. Rpλq is also bounded function on C with Rpλq Ñ 0 as λ Ñ 8.
Indeed, by Proposition 4.4.4, Rpλq is a bounded in the annulus |λ| ¥ 2}T } and
vanishes at infinity. Since Rpλq is a continuous function by Corollary 4.4.7, Rpλq is
also bounded in the disc |λ| ¤ 2}T }.
Claim. By Liouville’s theorem, Rpλq 0 everywhere.
Indeed, we fix a functional f P LpX, X q and apply the usual Liouville’s theo-
rem for the bounded entire function f pRpλqq. It follows that f pRpλqq is constant,
and since it must vanish at infinity it is zero everywhere. The claim follows.
The last claim contradicts the fact that Rpλq is an invertible operator.
Summarizing our findings, we can state that the spectrum of every bounded
linear operator is a nonempty compact subset of C.
4.4.4. Spectral radius. The spectrum of any operator T P LpX, X q is a
Lec. 35: 12/1
(i) f pRpλqq is an analytic function on the annulus |λ| ¡ rpT q by Corollary 4.4.4;
(ii) f pRpλqq is represented by a convergent Laurent series
8̧
(4.7) f pRpλqq λk1 f pT k q
k 1
in the smaller annulus |λ| ¡ }T }.
By the theory of convergence of Laurent series (outlined in the footnote), the
series (4.7) converges in the larger annulus |λ| ¡ rpT q. So the terms of the series
are bounded for such λ:
sup λn1 f pT n q 8.
n
This means that the sequence pλn1 T n ) is weakly bounded in LpX, X q. By a
consequence of the principle of uniform boundedness (Corollary 3.3.5), this sequence
is (strongly) bounded, i.e.
sup λn1 T n : K 8.
n
Taking n-th root and rearranging the terms, we obtain }T n }1{n ¤ K 1{n λ1 1{n for
all n. It follows that lim supn }T n }1{n ¤ |λ|. Since this happens for all λ such that
|λ| ¡ rpT q, we have proved that
lim sup }T n }1{n ¤ rpT q.
n
So, putting this together with the upper bound, we have proved that
rpT q ¤ inf }T n }1{n ¤ limninf }T n }1{n ¤ lim sup }T n }1{n ¤ rpT q.
n n
This completes the proof.
4.4.5. Spectrum of compact operators. As compact operators are proxies
of finite rank operators, one is able to fully classify their spectrum. First of all, for
every T P K pX, X q one has
0 P σ pT q
since T is not invertible by Corollary 4.1.8.
Exercise 4.4.14. Construct three examples of compact operators for
which 0 is in the point, continuous, and residual spectrum respectively.
There exist unique radii r, R P R X t8u such that the Laurent series converges in the annulus
A t z P C : r | z z0 | R u
and diverges outside the closure of A. Moreover, there exists at least one point on the inner
t P
boundary z C : z z0 | | u r of A and at least one point on the outer boundary z C : t P
| | u
z z0 pq
R of A such that f z can not be analytically continues to those points.
4.4. PROPERTIES OF SPECTRUM. SPECTRUM OF COMPACT OPERATORS. 107
Proof. Clearly, the second and third claims of the theorem follow from the
first one (why?). So, assume the contrary, that there exist ε ¡ 0 and an infinite
sequence of linearly independent vectors pxk q8
k1 such that
λk xk , where |λk | ¡ ε.
T xk
Consider the subspaces En Spanpxk qnk1 ; then E1 E2 is a sequence of
proper inclusions. Therefore we can choose vectors
Then
T yn λn apnnq xn vn1 , T un1 P En1 .
where vn1
Now we are ready to estimate }T yn T ym } for n ¡ m. Since T ym P Em En1 ,
we obtain
}T yn T ym } }λn apnnq xn wn1 } where wn1 P En1
}λn yn zn1 } where wn1 P En1
¥ |λn | distpyn , En1 q ¥ 2ε .
It follows that pT yn q8
n1 contains no Cauchy subsequences as claimed. The proof
is complete.
Proposition 4.4.16 (Classification of spectrum of compact operators). Let
T P K pX, X q be a compact operator on a Banach space X. Then
σ pT q σp pT q Y t0u.
Proof. As we already noticed in the beginning of this section, 0 P σ pT q. Let
now λ P σ pT q, λ 0. Fredholm alternative (Theorem 4.2.2) states that either
T λI is not injective (in which case λ P σp pT q) or T λI is both injective and
surjective. In the latter case T λI is invertible by the inverse mapping theorem,
which means that λ R σ pT q. The proof is complete.
4.4. PROPERTIES OF SPECTRUM. SPECTRUM OF COMPACT OPERATORS. 108
σ pU q tλ P C : |λ| 1u.
4.4.7. Additional exercises. In the following two exercises, one can work
over R. Similar results hold over C. The only difference is that for Hilbert spaces,
one has to take complex conjugation in appropriate places (which ones?), see Re-
mark 2.4.26.
Exercise 4.4.23. [Spectrum of adjoint I] Let T P LpX, X q. Prove that
σ pT q σ pT q. Here the bar stands for complex conjugation rather than
for closure.
x y xx, T xy xT x, xy.
1This follows from the identity T x, x
110
5.1. SPECTRUM OF SELF-ADJOINT OPERATORS 111
(5.1) xT x, yy 41 f px y q f px y q if px iy q if px iy q
This is a slight generalization of the polarization identity of Proposition 1.4.19,
where T I. (Prove it!)
The norm of T can be conveniently computed from the quadratic form f pT q:
Proposition 5.1.5 (Norm of a self-adjoint operator). For every self-adjoint
operator T P LpH, H q, one has
}T } sup |xT x, xy|.
P
x SH
It remains to show that the inequality here is actually the identity. To this end, we
note that
sup |xT x, y y| sup RexT x, y y
P
x,y SH P
x,y SH
and use the real part of polarization identity (5.1):
1
RexT x, y y xT px y q, x y y xT px y q, x y y
4
¤ M4 }x y}2 }x y}2 (by the definition of M )
¤ M4 2}x}2 2}x}2 (by the parallelogram law)
¤ M (as }x} }y} 1).
This completes the proof.
5.1.3. Criterion of spectrum points. We would like to study the spectrum
of self-adjoint operators T P LpH, H q. An easy observation is that all eigenvalues
of T must be real, that is
σp pT q R.
Indeed, if λ is an eigenvalue with an eigenvector x then xT x, xy xλx, xy λxx, xy
which must be the same as xx, T xy xx, λxy λxx, xy. This shows that λ λ, so
λ P R.
We will soon prove that the whole spectrum of T is real, i.e. σ pT q R, and
moreover we will compute the smallest interval containing σ pT q. Let us start with
ruling out the residual spectrum:
Proposition 5.1.6 (No residual spectrum). Let T P LpH, H q be a self-adjoint
operator. Then σr pT q H.
Proof. Let λ P σr pT q. This means that kerpT λI q 0 while ImpT λI q is
not dense in H. Since λ is not an eigenvalue, λ is not an eigenvalue either (recall
that all eigenvalues of T must be real). Using this and the duality relation (2.4.32),
we obtain that pImpT λI qqK kerpT λI q kerpT λI q 0. It follows that
ImpT λI q is dense in H, which is a contradiction.
5.1. SPECTRUM OF SELF-ADJOINT OPERATORS 112
We see that T now has a quite simple form, which we studied in Example 4.3.5.
In literature, one comes across various forms of spectral Theorem 5.2.5. We
mention two of them. Let as before pφk q denote an orthonormal basis of eigenvectors
of T with corresponding eigenvalues λk . Orthogonal basis expansion gives
¸
x xx, φk yφk , x P H.
k
Applying operator T for both sides and using that T φk λk φk we obtain that
¸
Tx λn xx, φn yφn , x P H.
n
³1
Proof. Consider the integral operator pT f qptq 0 k pt, sqf psq ds on L2 r0, 1s.
Let pφn q be an orthonormal basis of its eigenvectors. Then the functions
φn ptqφm psq, n, m 1, 2, . . .
ψnm
form an orthonormal basis of L2 pr0, 1s2 q. (Check!)
Let us write the basis expansion of our function in L2 pr0, 1s2 q:
¸
k xk, ψnm yψnm .
n,m
The spectrum of a polynomial ppT q can be easily computed from the spectrum
of T :
Theorem 5.3.9 (Spectral mapping theorem). Let pptq be a polynomial and
T P LpH, H q. Then4
σ pppT qq ppσ pT qq.
Proof. For every complex number λ, we have λ P σ pppT qq if and only if the
operator ppT q λI pp λqpT q is not invertible. By the invertibility Lemma 5.3.8,
this is equivalent to the condition that pp λqptq 0 for some t P σ pT q, which
means that pptq λ for some t P σ pT q. The latter is equivalent to λ P ppσ pT qq.
Using the spectral mapping theorem, one can in particular easily compute the
norms of operator polynomials:
Corollary 5.3.10 (Operator norm of polynomials). Let pptq be a polynomial
and T P LpH, H q be a self-adjoint operator. Then
}ppT q} tmax
PσpT q
|pptq|.
This result generalizes the identity rpT q }T } for the spectral radius of self-
adjoint operators T proved in Corollary 5.1.11.
Proof. Let us apply Corollary 5.1.11 for the operator ppT q. Then spectral
mapping theorem yields
}ppT q} rpppT qq tPσmax
pppT qq
|t| tPpmax
pσpT qq
|t| smax
PσpT q
|ppsq|
as claimed.
(ii) Since the operators pn pT q are self-adjoint, and the self-adjoint operators
form a closed subset of LpH, H q (Exercise 5.1.4), f pT q is also self-adjoint. Further-
more, repeating the estimate in part (i), one sees that for any other approximating
sequence of polynomials qn ptq one has }pn pT q qn pT q} Ñ 0 as n Ñ 8. It follows
that the limit f pT q must be the same whether one chooses pn pT q or qn pT q as an
approximating sequence.
5Formally, since Weierstrass theorem is typically stated for functions on an interval, we first
pq
extend f t to a continuous function on some interval m, M r s
p q
σ T (for example, the spectral
interval). Such extension can be done e.g. by Tietze extension theorem.
5.3. POSITIVE OPERATORS. CONTINUOUS FUNCTIONAL CALCULUS 120
? self-adjoint operator
5.3.6. Square root of an operator. Consider a positive
T P LpH, H q. Then σ pT q?
r0, 8q. The function f ptq t is continuous on r0, 8q,
? 2 f pT q T .
so we can define ?
? tq t, the algebra homomorphism
Since p ? property implies that p T q2 T .
Since t ¥ 0, Corollary 5.3.15 implies that T is a positive self-adjoint operator.
Summarizing, we have proved the following (except uniqueness):
Proposition 5.3.18 (Square root of an operator). For every positive self-
?
adjoint operator T P LpH, H q, there exists a unique positive self-adjoint operator
T P LpH, H q such that ?
p T q2 T.
5.3. POSITIVE OPERATORS. CONTINUOUS FUNCTIONAL CALCULUS 121
?
Exercise 5.3.19. Prove the uniqueness of T.
As an application of operator square root, we will now prove the following
result, whose formulation has nothing to do with functional calculus
Proposition 5.3.20. Let S, T P LpH, H q be a pair of commuting self-adjoint
operators. If S ¥ 0 and T ¥ 0 then ST ¥ 0.
For compact operators S, T this would follow because in this case one can find
a common basis of eigenvectors in H. (Do this!) In general, I don’t know any proof
that won’t involve functional calculus.
?
? Proof. We want to show that xST x, xy ¥ 0 for every x P H.?Since ?S ? and
T are communing self-adjoint operators (see Exercise 5.3.17), and ST S T .
(Why?) So we have
? ? ? ? ? ? ? ? ? ? ? ?
xST x, xy x S S T T x, xy x T S S T x, xy x S T x, S T xy ¥ 0
as required.
Exercise 5.4.3. Compute the spectral measures for the diagonal ma-
trix T diagpλ1 , . . . , λn q acting as an operator on Cn .
(ii) For every decomposition rm, M s 8k1 Ek into disjoint sets Borel Ek , one
has
8̧
I PE n ,
n 1
where the pointwise convergence of the series.
These properties are strikingly similar to the axioms of a Borel probability
measure on rm, M s. However, the usual probability measures are functions λ : E ÞÑ
λpE q with values in r0, 1s, while PE takes values in LpH, H q, more specifically in the
operator interval between 0 and I. For this reason, the assignment Pλ : E ÞÑ PE
for Borel subsets E rm, M s is called a projection-valued measure for T .
The projection-valued measure can be reduced to the usual measure in a simple
way. For x P H, the assignment E ÞÑ xPE x, xy is clearly a usual Borel measure
on rm, M s. This follows from Proposition 5.4.6. Moreover, (5.5) shows that this is
nothing else than the spectral measure µx,x :
»M
xPλ pE qx, xy xPE x, xy 1E ptq dµx,x pλq µx,x pE q.
m
For this reason, the projection-valued measure Pλ itself, rather than µx,x , is often
called the spectral measure associated with the operator T .
5.4.4. Spectral theorem for self-adjoint operators.
Theorem 5.4.7 (Spectral theorem). Let T P LpH, H q be a self-adjoint operator
on a Hilbert space H. Then » 8
T λ dPλ
8
where Pλ is the projection-valued measure associated with T . The integral in fact
can be taken over a finite interval rm, M s containing the spectrum of T .
The integration with respect to the operator-valued measure Pλ is understood
in the sense that » M
xT x, xy λ xdPλ x, xy;
m
As we noted, xdPλ x, xy is just the spectral measure µx,x , so the last integral is the
usual Lebesgue integral.
Proof. With this remark, Theorem 5.4.7 is a reformulation of a partial case
of (5.6):
»M
xT x, xy λ dµx,x pλq.
m
Theorem 5.4.7 should be compared to the spectral Theorem 5.3 for compact
self-adjoint operators T . According to this theorem, T can be decomposed into the
sum ¸
T λ k Pk
k
where λk are the eigenvalues of T and Pk denotes the orthogonal projection in H
onto the eigenspace corresponding to the eigenvalue λk , see (5.3).
As we know, for general operators T (not necessarily compact), the spectrum
of T may no longer be countable. So the sum is replaced by the integral in Theo-
rem 5.4.7.
Bibliography
[1] K. Ball, An elementary introduction to modern convex geometry. Flavors of geometry, 1–58,
Math. Sci. Res. Inst. Publ., 31, Cambridge Univ. Press, Cambridge, 1997.
[2] L. Carleson, On convergence and growth of partial sums of Fourier series, Acta Mathe-
matica 116 (1966), 135–157.
[3] P. Enflo, A counterexample to the approximation problem in Banach spaces., Acta Math.
130 (1973), 309–317.
[4] E. Gluskin, The diameter of the Minkowski compactum is roughly equal to n, Funktsional.
Anal. i Prilozhen. 15 (1981), 72–73.
[5] W. T. Gowers, B. Maurey, The unconditional basic sequence problem, J. Amer. Math. Soc.
6 (1993), 851–874
[6] J. Lindenstrauss, Joram, On complemented subspaces of m, Israel J. Math. 5 (1967), 153–
156.
[7] J. Lindenstrauss, L. Tzafriri, On the complemented subspaces problem, Israel J. Math. 9
(1971), 263–269.
[8] J. Lindenstrauss, L. Tzafriri, Classical Banach spaces. I and II. Springer-Verlag, , 1977,
1979.
[9] A. Pelczkynski, Projections in certain Banach spaces, Studia Math. 19 (1960), 209–228.
[10] R. S. Phillips, On linear transformations, Trans. Amer. Math. Soc. 48 (1940), 516–541.
126
Index
127
INDEX 128
Schauder basis, 79
Schauder system, 80
Schur property, 89
Self-adjoint operator, 110
Semi-norm, 14
Separable space, 35