A First Course in Functional Analysis Theory and Applications
A First Course in Functional Analysis Theory and Applications
VS2%RS]BZRS%qS2Zqx%BqphSVqphiR%R
nYBiSpqHSV{{h%xp%BqR
gpA%qHpqpYS6q
VqY7S^RR
VqS%7{%qSBPSF%7AhHBqS^ZAh%RY%qS]B7{pqi
www:anthempress:com
nY%RSH%%BqSfRS{ZAh%RYHS%qS<~SpqHS<6VS0(ck
AiSVWn;dS^gd66
}ur}DS!hpx,P%pRSgBpHSNBqHBqS6dcSM;VS<~
BS^"S!B-S}}SNBqHBqS6FcS}mS<~
pqH
0::SpH%RBqSV&S'ccDSWaSOB,SWOSc((cDS<6V
S]B{i%YS©SgpA%qHpqpYS6qS0(ck
nYSpZYBSpRRRSYS7BphS%YSBSAS%Hq%fHSpRSYSpZYBSBPSY%RSaB,
VhhS%YRSR&HSF%YBZSh%7%%qSYS%YRSZqHSxB{i%YSR&HSpAB&
qBS{pSBPSY%RS{ZAh%xp%BqS7piSAS{BHZxHSRBHSBS%qBHZxHS%qB
pS%&phSRiR7SBSpqR7%HS%qSpqiSPB7SBSAiSpqiS7pqR
thxBq%xS7xYpq%xphS{YBBxB{i%qSxBH%qSBSBYa%RCS
a%YBZSYS{%BSa%qS{7%RR%BqSBPSABYSYSxB{i%YS
BaqSpqHSYSpAB&S{ZAh%RYSBPSY%RSABB,
BritishLibraryCataloguingTinTPublicationData
VSxpphBZSxBHSPBSY%RSABB,S%RSp&p%hpAhSPB7SYS!%%RYSN%Api
LibraryofCongressCatalogingTinTPublicationData
VSxpphBSxBHSPBSY%RSABB,SYpRSAqSJZRH
v6!WTckUS}MS(SMu}0MSk0(S}St;A,C
v6!WTc(US(SMu}0MSk0(S(St;A,C
nY%RS%hS%RSphRBSp&p%hpAhSpRSpqS!BB,
Preface
This book is the outgrowth of the lectures delivered on functional
analysis and allied topics to the postgraduate classes in the Department
of Applied Mathematics, Calcutta University, India. I feel I owe an
explanation as to why I should write a new book, when a large number of
books on functional analysis at the elementary level are available. Behind
every abstract thought there is a concrete structure. I have tried to unveil
the motivation behind every important development of the subject matter.
I have endeavoured to make the presentation lucid and simple so that the
learner can read without outside help.
The first chapter, entitled ‘Preliminaries’, contains discussions on topics
of which knowledge will be necessary for reading the later chapters. The
first concepts introduced are those of a set, the cardinal number, the
different operations on a set and a partially ordered set respectively.
Important notions like Zorn’s lemma, Zermelo’s axiom of choice are stated
next. The concepts of a function and mappings of different types are
introduced and exhibited with examples. Next comes the notion of a linear
space and examples of different types of linear spaces. The definition of
subspace and the notion of linear dependence or independence of members
of a subspace are introduced. Ideas of partition of a space as a direct
sum of subspaces and quotient space are explained. ‘Metric space’ as an
abstraction of real line 4 is introduced. A broad overview of a metric
space including the notions of convergence of a sequence, completeness,
compactness and criterion for compactness in a metric space is provided in
the first chapter. Examples of a non-metrizable space and an incomplete
metric space are also given. The contraction mapping principle and its
application in solving different types of equations are demonstrated. The
concepts of an open set, a closed set and an neighbourhood in a metric
space are also explained in this chapter. The necessity for the introduction
of ‘topology’ is explained first. Next, the axioms of a topological space are
stated. It is pointed out that the conclusions of the Heine-Borel theorem
in a real space are taken as the axioms of an abstract topological space.
Next the ideas of openness and closedness of a set, the neighbourhood of
a point in a set, the continuity of a mapping, compactness, criterion for
compactness and separability of a space naturally follow.
Chapter 2 is entitled ‘Normed Linear Space’. If a linear space admits a
metric structure it is called a metric linear space. A normed linear space is a
type of metric linear space, and for every element x of the space there exists
a positive number called norm x or x fulfilling certain axioms. A normed
linear space can always be reduced to a metric space by the choice of a
suitable metric. Ideas of convergence in norm and completeness of a normed
linear space are introduced with examples of several normed linear spaces,
Banach spaces (complete normed linear spaces) and incomplete normed
linear spaces.
vii
Continuity of a norm and equivalence of norms in a finite dimensional
normed linear space are established. The definition of a subspace and its
various properties as induced by the normed linear space of which this
is a subspace are discussed. The notion of a quotient space and its role
in generating new Banach spaces are explained. Riesz’s lemma is also
discussed.
Chapter 3 dwells on Hilbert space. The concepts of inner product space,
complete inner product or Hilbert space are introduced. Parallelogram law,
orthogonality of vectors, the Cauchy-Bunyakovsky-Schwartz inequality, and
continuity of scalar (inner) product in a Hilbert space are discussed. The
notions of a subspace, orthogonal complement and direct sum in the setting
of a Hilbert space are introduced. The orthogonal projection theorem takes
a special place.
Orthogonality, various orthonormal polynomials and Fourier series are
discussed elaborately. Isomorphism between separable Hilbert spaces is
also addressed. Linear operators and their elementary properties, space
of linear operators, linear operators in normed linear spaces and the norm
of an operator are discussed in Chapter 4. Linear functionals, space of
bounded linear operators and the uniform boundedness principle and its
applications, uniform and pointwise convergence of operators and inverse
operators and the related theories are presented in this chapter. Various
types of linear operators are illustrated. In the next chapter, the theory of
linear functionals is discussed. In this chapter I introduce the notions of
a linear functional, a bounded linear functional and the limiting process,
and assert continuity in the case of boundedness of the linear functional
and vice-versa. In the case of linear functionals apart from different
examples of linear functionals, representation of functionals in different
Banach and Hilbert spaces are studied. The famous Hahn-Banach theorem
on the extension on a functional from a subspace to the entire space with
preservation of norm is explained and the consequences of the theorem
are presented in a separate chapter. The notions of adjoint operators and
conjugate space are also discussed. Chapter 6 is entitled ‘Space of Bounded
Linear Functionals’. The chapter dwells on the duality between a normed
linear space and the space of all bounded linear functionals on it. Initially
the notions of dual of a normed linear space and the transpose of a bounded
linear operator on it are introduced. The zero spaces and range spaces of a
bounded linear operator and of its duals are related. The duals of Lp ([a, b])
and C([a, b]) are described. Weak convergence in a normed linear space
and its dual is also discussed. A reflexive normed linear space is one for
which the canonical embedding in the second dual is surjective (one-to-
one). An elementary proof of Eberlein’s theorem is presented. Chapter 7 is
entitled ‘Closed Graph Theorem and its Consequences’. At the outset the
definitions of a closed operator and the graph of an operator are given. The
closed graph theorem, which establishes the conditions under which a closed
linear operator is bounded, is provided. After introducing the concept of an
viii
open mapping, the open mapping theorem and the bounded inverse theorem
are proved. Application of the open mapping theorem is also provided. The
next chapter bears the title ‘Compact Operators on Normed Linear Spaces’.
Compact linear operators are very important in applications. They play a
crucial role in the theory of integral equations and in various problems of
mathematical physics. Starting from the definition of compact operators,
the criterion for compactness of a linear operator with a finite dimensional
domain or range in a normed linear space and other results regarding
compact linear operators are established. The spectral properties of a
compact linear operator are studied. The notion of the Fredholm alternative
is discussed and the relevant theorems are provided. Methods of finding an
approximate solution of certain equations involving compact operators in
a normed linear space are explored. Chapter 9 bears the title ‘Elements of
Spectral Theory on Self-adjoint Operators in Hilbert Spaces’. Starting from
the definition of adjoint operators, self-adjoint operators and their various
properties are elaborated upon the context of a Hilbert space. Quadratic
forms and quadratic Hermitian forms are introduced in a Hilbert space and
their bounds are discovered. I define a unitary operator in a Hilbert space
and the situation when two operators are said to be unitarily equivalent,
is explained. The notion of a projection operator in a Hilbert space is
introduced and its various properties are investigated. Positive operators
and the square root of operators in a Hilbert space are introduced and
their properties are studied. The spectrum of a self-adjoint operator in a
Hilbert space is studied and the point spectrum and continuous spectrum
are explained. The notion of invariant subspaces in a Hilbert space is
also brought within the purview of the discussion. Chapter 10 is entitled
‘Measure and Integration in Spaces’. In this chapter I discuss the theory
4
of Lebesgue integration and p-integrable functions on . Spaces of these
functions provide very useful examples of many theorems in functional
analysis. It is pointed out that the concept of the Lebesgue measure is
a generalization of the idea of subintervals of given length in 4 to a class
4
of subsets in . The ideas of the Lebesgue outer measure of a set E ⊂ , 4
Lebesgue measurable set E and the Lebesgue measure of E are introduced.
The notions of measurable functions and integrable functions in the sense
of Lebesgue are explained. Fundamental theorems of Riemann integration
and Lebesgue integration, Fubini and Toneli’s theorem, are stated and
explained. Lp spaces (the space of functions p-integrable on a measure
4
subset E of ) are introduced, that (E) is complete and related properties
discussed. Fourier series and then Fourier integral for functions are
investigated. In the next chapter, entitled ‘Unbounded Linear Operators’,
I first give some examples of differential operators that are not bounded.
But these are closed operators, or at least have closed linear extensions. It
is indicated in this chapter that many of the important theorems that hold
for continuous linear operators on a Banach space also hold for closed linear
operators. I define the different states of an operator depending on whether
ix
the range of the operator is the whole of a Banach space or the closure of
the range is the whole space or the closure of the range is not equal to
the space. Next the characterization of states of operators is presented.
Strictly singular operators are then defined and accompanied by examples.
Operators that appear in connection with the study of quantum mechanics
also come within the purview of the discussion. The relationship between
strictly singular and compact operators is explored. Next comes the study
of perturbation theory. The reader is given an operator ‘A’, the certain
properties of which need be found out. If ‘A’ is a complicated operator, we
sometimes express ‘A = T +B’ where ‘T ’ is a relatively simple operator and
‘B’ is related to ‘T ’ in such a manner that knowledge about the properties
of ‘T ’ is sufficient to gain information about the corresponding properties
of ‘A’. In that case, for knowing the specific properties of ‘A’, we can
replace ‘A’ with ‘T ’, or in other words we can perturb ‘A’ by ‘T ’. Here
we study perturbation by a bounded linear operator and perturbation by
strictly singular operator. Chapter 12 bears the title ‘The Hahn-Banach
Theorem and the Optimization Problems’. I first explain an optimization
problem. I define a hyperplane and describe what is meant by separating
a set into two parts by a hyperplane. Next the separation theorems for
a convex set are proved with the help of the Hahn-Banach theorem. A
minimum Norm problem is posed and the Hahn-Banach theorem is applied
to the proving of various duality theorems. Said theorem is applied to prove
Chebyshev approximation theorems. The optimal control problem is posed
and the Pontryagin’s problem is mentioned. Theorems on optimal control of
rockets are proved using the Hahn-Banach theorem. Chapter 13 is entitled
‘Variational Problems’ and begins by introducing a variational problem.
The aim is to investigate under which conditions a given functional in a
normed linear space admits of an optimum. Many differential equations are
often difficult to solve. In such cases a functional is built out of the given
equation and minimized. One needs to show that such a minimum solves
the given equation. To study those problems, a Gâteaux derivative and a
Fréchet derivative are defined as a prerequisite. The equivalence of solving
a variational problem and solving a variational inequality is established.
I then introduce the Sobolev space to study the solvability of differential
equations. In Chapter 14, entitled ‘The Wavelet Analysis’, I provide a
brief introduction to the origin of wavelet analysis. It is the outcome of
the confluence of mathematics, engineering and computer science. Wavelet
analysis has begun to play a serious role in a broad range of applications
including signal processing, data and image compression, the solving of
partial differential equations, the modeling of multiscale phenomena and
statistics. Starting from the notion of information, we discuss the scalable
structure of information. Next we discuss the algebra and geometry of
wavelet matrices like Haar matrices and Daubechies’s matrices of different
ranks. Thereafter come the one-dimensional wavelet systems where the
scaling equation associated with a wavelet matrix, the expansion of a
x
function in terms of wavelet system associated with a matrix and other
results are presented. The final chapter is concerned with dynamical
systems. The theory of dynamical systems has its roots in the theory of
ordinary differential equations. Henry Poincaré and later Ivar Benedixon
studied the topological properties of the solutions of autonomous ordinary
differential equations (ODEs) in the plane. They did so with a view of
studying the basic properties of autonomous ODEs without trying to find
out the solutions of the equations. The discussion is confined to one-
dimensional flow only.
Prerequisites The reader of the book is expected to have a knowledge
of set theory, elements of linear algebra as well as having been exposed to
metric spaces.
Courses The book can be used to teach two semester courses at the M.Sc.
level in universities (MS level in Engineering Institutes):
(i) Basic course on functional analysis. For this Chapters 2–9 may be
consulted.
(ii) Another course may be developed on linear operator theory. For
this Chapters 2, 3–5, 7–9 and 11 may be consulted. The Lebesgue
measure is discussed at an elementary level in Chapter 10; Chapters
2–9 can, however, be read without any knowledge of the Lebesgue
measure.
xi
>=C4=CB
$DJHE:K9J?ED [YLL
$ *H;B?C?D7H?;I ^
7HW
*XQFWLRQ 1DSSLQJ
0LQHDU 7SDFH
1HWULF 7SDFHV
8RSRORJLFDO 7SDFHV
'RQWLQXLW\ 'RPSDFWQHVV
$$ (EHC;: &?D;7H -F79;I ^
(HQLWLRQV DQG )OHPHQWDU\ 4URSHUWLHV
7XEVSDFH 'ORVHG 7XEVSDFH
*LQLWH (LPHQVLRQDO 2RUPHG 0LQHDU 7SDFHV DQG 7XEVSDFHV
5XRWLHQW 7SDFHV
'RPSOHWLRQ RI 2RUPHG 7SDFHV
$$$ #?B8;HJ -F79; ^
-QQHU 4URGXFW 7SDFH ,LOEHUW 7SDFH
]pZxYiT!Zqip,B&R,iT6xYap8St]!6CSvqJZph%i
'DXFK\&XQ\DNRYVN\7FKZDUW] -QHTXDOLW\
4DUDOOHORJUDP 0DZ
3UWKRJRQDOLW\
3UWKRJRQDO 4URMHFWLRQ 8KHRUHP
3UWKRJRQDO 'RPSOHPHQWV (LUHFW 7XP
3UWKRJRQDO 7\VWHP
'RPSOHWH 3UWKRQRUPDO 7\VWHP
-VRPRUSKLVP EHWZHHQ 7HSDUDEOH ,LOEHUW 7SDFHV
$0 &?D;7H )F;H7JEHI ^
(HQLWLRQ! 0LQHDU 3SHUDWRU
0LQHDU 3SHUDWRUV LQ 2RUPHG 0LQHDU 7SDFHV
0LQHDU *XQFWLRQDOV
8KH 7SDFH RI &RXQGHG 0LQHDU 3SHUDWRUV
9QLIRUP &RXQGHGQHVV 4ULQFLSOH
7RPH %SSOLFDWLRQV
-QYHUVH 3SHUDWRUV
&DQDFK 7SDFH ZLWK D &DVLV
[LLL
0 &?D;7H !KD9J?ED7BI ^
,DKQ&DQDFK 8KHRUHP
,DKQ&DQDFK 8KHRUHP IRU 'RPSOH[ :HFWRU DQG 2RUPHG
0LQHDU 7SDFH
%SSOLFDWLRQ WR &RXQGHG 0LQHDU *XQFWLRQDOV RQ ?8 9@
8KH +HQHUDO *RUP RI 0LQHDU *XQFWLRQDOV LQ 'HUWDLQ
*XQFWLRQDO 7SDFHV
8KH +HQHUDO *RUP RI 0LQHDU *XQFWLRQDOV LQ ,LOEHUW
7SDFHV
'RQMXJDWH 7SDFHV DQG %GMRLQW 3SHUDWRUV
0$ -F79; E< EKD:;: &?D;7H !KD9J?ED7BI ^
'RQMXJDWHV (XDOV DQG 8UDQVSRVHV %GMRLQWV
'RQMXJDWHV (XDOV RI (9 ?8 9@ DQG ?8 9@
;HDN DQG ;HDN 'RQYHUJHQFH
6H H[LYLW\
&HVW %SSUR[LPDWLRQ LQ 6H H[LYH 7SDFHV
0$$ BEI;: "H7F> .>;EH;C 7D: $JI EDI;GK;D9;I ^
'ORVHG +UDSK 8KHRUHP
3SHQ 1DSSLQJ 8KHRUHP
&RXQGHG -QYHUVH 8KHRUHP
%SSOLFDWLRQ RI WKH 3SHQ 1DSSLQJ 8KHRUHP
V{{h%xp%BqRSBPSYS"{qSp{{%qSnYB7
0$$$ ECF79J )F;H7JEHI ED (EHC;: &?D;7H -F79;I
'RPSDFW 0LQHDU 3SHUDWRUV
7SHFWUXP RI D 'RPSDFW 3SHUDWRU
*UHGKROP %OWHUQDWLYH
%SSUR[LPDWH 7ROXWLRQV
$2 B;C;DJI E< -F;9?7B .>;EHO E< -;B<
:@E?DJ
dh7qRSBPS6{xphSnYBiSBPS6hPTVH_B%q ^
)F;H7JEHI ?D #?B8;HJ -F79;I
"{pBRS%qS;%hAS6{pxR
%GMRLQW 3SHUDWRUV
7HOI%GMRLQW 3SHUDWRUV
5XDGUDWLF *RUP
9QLWDU\ 3SHUDWRUV 4URMHFWLRQ 3SHUDWRUV
4RVLWLYH 3SHUDWRUV 7TXDUH 6RRWV RI D 4RVLWLYH 3SHUDWRU
7SHFWUXP RI 7HOI%GMRLQW 3SHUDWRUV
-QYDULDQW 7XEVSDFHV
'RQWLQXRXV 7SHFWUD DQG 4RLQW 7SHFWUD
[LY
2 ';7IKH; 7D: $DJ;=H7J?ED ?D -F79;I ^
8KH 0HEHVJXH 1HDVXUH RQ
1HDVXUDEOH DQG 7LPSOH *XQFWLRQV
'DOFXOXV ZLWK WKH 0HEHVJXH 1HDVXUH
8KH *XQGDPHQWDO 8KHRUHP IRU 6LHPDQQ -QWHJUDWLRQ
8KH *XQGDPHQWDO 8KHRUHP IRU 0HEHVJXH -QWHJUDWLRQ
(9 7SDFHV DQG 'RPSOHWHQHVV
(9 'RQYHUJHQFH RI *RXULHU 7HULHV
2$ /D8EKD:;: &?D;7H )F;H7JEHI ^
(HQLWLRQ! %Q 9QERXQGHG 0LQHDU 3SHUDWRU
7WDWHV RI D 0LQHDU 3SHUDWRU
(HQLWLRQ! 7WULFWO\ 7LQJXODU 3SHUDWRUV
6HODWLRQVKLS &HWZHHQ 7LQJXODU DQG 'RPSDFW 3SHUDWRUV
4HUWXUEDWLRQ E\ &RXQGHG 3SHUDWRUV
4HUWXUEDWLRQ E\ 7WULFWO\ 7LQJXODU 3SHUDWRUV
4HUWXUEDWLRQ LQ D ,LOEHUW 7SDFH DQG %SSOLFDWLRQV
2$$ .>; #7>D
7D79> .>;EH;C 7D: )FJ?C?P7J?ED ^
*HE8B;CI
8KH 7HSDUDWLRQ RI D 'RQYH[ 7HW
1LQLPXP 2RUP 4UREOHP DQG WKH (XDOLW\ 8KHRU\
%SSOLFDWLRQ WR 'KHE\VKHY %SSUR[LPDWLRQ
%SSOLFDWLRQ WR 3SWLPDO 'RQWURO 4UREOHPV
2$$$ 07H?7J?ED7B *HE8B;CI ^
1LQLPL]DWLRQ RI *XQFWLRQDOV LQ D 2RUPHG 0LQHDU 7SDFH
+ADWHDX[ (HULYDWLYH
*UHFKHW (HULYDWLYH
dJZ%&phqxSBPSYS%q%7%8p%BqS^BAh7SPBS6Bh&%q
)TXLYDOHQFH RI WKH 1LQLPL]DWLRQ 4UREOHP WR 7ROYLQJ
:DULDWLRQDO -QHTXDOLW\
wp%p%BqphSvqJZph%i
(LVWULEXWLRQV
7REROHY 7SDFH
2$0 .>; 17L;B;J D7BOI?I ^
%Q -QWURGXFWLRQ WR ;DYHOHW %QDO\VLV
8KH 7FDODEOH 7WUXFWXUH RI -QIRUPDWLRQ
%OJHEUD DQG +HRPHWU\ RI ;DYHOHW 1DWULFHV
3QHGLPHQVLRQDO ;DYHOHW 7\VWHPV
"qT*%7qR%BqphSFp&hS6iR7R
[Y
20 OD7C?97B -OIJ;CI ^
% (\QDPLFDO 7\VWHP DQG -WV 4URSHUWLHV
,RPHRPRUSKLVP (LHRPRUSKLVP 6LHPDQQLDQ
1DQLIROG
7WDEOH 4RLQWV 4HULRGLF 4RLQWV DQG 'ULWLFDO 4RLQWV
)[LVWHQFH 9QLTXHQHVV DQG 8RSRORJLFDO 'RQVHTXHQFH
d-%RqxS<q%JZqRRSpqHSnB{BhB%xphS]BqRJZqxR
&LIXUFDWLRQ 4RLQWV DQG 7RPH 6HVXOWV
&?IJ E< -OC8EBI ^
?8B?E=H7F>O ^
$D:;N ^
[YL
Introduction
Functional analysis is an abstract branch of mathematics that grew
out of classical analysis. It represents one of the most important
branches of the mathematical sciences. Together with abstract algebra and
mathematical analysis, it serves as a foundation of many other branches of
mathematics. Functional analysis is in particular widely used in probability
and random function theory, numerical analysis, mathematical physics and
their numerous applications. It serves as a powerful tool in modern control
and information sciences.
The development of the subject started from the beginning of the
twentieth century, mainly through the initiative of the Russian school
of mathematicians. The impetus came from the developments of linear
algebra, linear ordinary and partial differential equations, calculus of
variation, approximation theory and, in particular, those of linear integral
equations, the theory of which had the greatest impact on the development
and promotion of modern ideas. Mathematicians observed that problems
from different fields often possess related features and properties. This
allowed for an effective unifying approach towards the problems, the
unification being obtained by the omission of inessential details. Hence
the advantage of such an abstract approach is that it concentrates on the
essential facts, so that they become clearly visible.
Since any such abstract system will in general have concrete realisations
(concrete models), we see that the abstract method is quite versatile in
its applications to concrete situations. In the abstract approach, one
usually starts from a set of elements satisfying certain axioms. The nature
of the elements is left unspecified. The theory then consists of logical
consequences, which result from the axioms and are derived as theorems
once and for all. This means that in the axiomatic fashion one obtains a
mathematical structure with a theory that is developed in an abstract way.
For example, in algebra this approach is used in connection with fields,
rings and groups. In functional analysis, we use it in connection with
‘abstract’ spaces; these are all of basic importance.
In functional analysis, the concept of space is used in a very wide and
surprisingly general sense. An abstract space will be a set of (unspecified)
elements satisfying certain axioms, and by choosing different sets of axioms,
we obtain different types of abstract spaces.
The idea of using abstract spaces in a systematic fashion goes back to M.
Fréchet (1906) and is justified by its great success. With the introduction
of abstract space in functional analysis, the language of geometry entered
the arena of the problems of analysis. The result is that some problems of
analysis were subjected to geometric interpretations. Furthermore many
conjectures in mechanics and physics were suggested, keeping in mind
the two-dimensional geometry. The geometric methods of proof of many
theorems came into frequent use.
xvii
The generalisation of algebraic concepts took place side by side with that
of geometric concepts. The classical analysis, fortified with geometric and
algebraic concepts, became versatile and ready to cope with new problems
not only of mathematics but also of mathematical physics. Thus functional
analysis should form an indispensable part of the mathematics curricula at
the college level.
xviii
CHAPTER 1
PRELIMINARIES
1.1 Set
The theory of sets is one of the principal tools of mathematics. One type of
study of set theory addresses the realm of logic, philosophy and foundations
of mathematics. The other study goes into the highlands of mathematics,
where set theory is used as a medium of expression for various concepts
in mathematics. We assume that the sets are ‘not too big’ to avoid any
unnecessary contradiction. In this connection one can recall the famous
‘Russell’s Paradox’ (Russell, 1959). A set is a collection of distinct and
distinguishable objects. The objects that belong to a set are called elements,
members or points of the set. If an object a belongs to a set A, then we
write a ∈ A. On the other hand, if a does not belong to A, we write
a∈ / A. A set may be described by listing the elements and enclosing them
in braces. For example, the set A formed out of the letters a, a, a, b, b, c can
be expressed as A = {a, b, c}. A set can also be described by some defining
properties. For example, the set of natural numbers can be written as
N = {x : x, a natural number} or {x|x, a natural number}. Next we
discuss set inclusion. If every element of a set A is an element of the set
B, A is said to be a subset of the set B or B is said to be a superset of
A, and this is denoted by A ⊆ B or B ⊇ A. Two sets A and B are said
to be equal if every element of A is an element of B and every element of
B is an element of A–in other words if A ⊆ B and B ⊆ A. If A is equal
to B, then we write A = B. A set is generally completely determined by
its elements, but there may be a set that has no element in it. Such a set
is called an empty (or void or null) set and the empty set is denoted by Φ
1
2 A First Course in Functional Analysis
(Phi). Φ ⊂ A; in other words, the null set is included in any set A – this
fact is vacuously satisfied. Furthermore, if A is a subset of B, A = Φ and
A = B, then A is said to be a proper subset of B (or B is said to properly
contain A). The fact that A is a proper subset of B is expressed as A ⊂ B.
Let A be a set. Then the set of all subsets of A is called the power set
of A and is denoted by P (A). If A has three elements like letters p, q and
r, then the set of all subsets of A has 8(= 23 ) elements. It may be noted
that the null set is also a subset of A. A set is called a finite set if it is
empty or it has n elements for some positive integer n; otherwise it is said
to be infinite. It is clear that the empty set and the set A are members of
P (A). A set A is called denumerable or enumerable if it is in one-to-one
correspondence with the set of natural numbers. A set is called countable
if it is either finite or denumerable. A set that is not countable is called
uncountable.
We now state without proof a few results which might be used in
subsequent chapters:
(i) An infinite set is equivalent to a subset of itself.
(ii) A subset of a countable set is a countable set.
The following are examples of countable sets: a) the set J of all integers,
b) the set 3 of all rational numbers, c) the set P of all polynomials with
rational coefficients, d) the set all straight lines in a plane each of which
passes through (at least) two different points with rational coordinates and
4
e) the set of all rational points in n .
Examples of uncountable sets are as follows: (i) an open interval ]a, b[, a
closed interval [a, b] where a = b, (ii) the set of all irrational numbers. (iii)
the set of all real numbers. (iv) the family of all subsets of a denumerable
set.
1.1.1 Cardinal numbers
Let all the sets be divided into two families such that two sets fall into
one family if and only if they are equivalent. This is possible because the
relation ∼ between the sets is an equivalence relation. To every such family
of sets, we assign some arbitrary symbol and call it the cardinal number of
each set of the given family. If the cardinal number of a set A is α, A = α
or card A = α. The cardinal number of the empty set is defined to be
0 (zero). We designate the number of elements of a nonempty finite set
as the cardinal number of the finite set. We assign ℵ0 to the class of all
denumerable sets and as such ℵ0 is the cardinal number of a denumerable
set. c, the first letter of the word ‘continuum’ stands for the cardinal number
of the set [0, 1].
1.1.2 The algebra of sets
In the following section we discuss some operations that can be
performed on sets. By universal set we mean a set that contains all the sets
Preliminaries 3
(a) x + y = y + x (commutativity);
(b) x + (y + z) = (x + y) + z (associativity);
(c) There exists a uniquely defined element 0, such that x + θ = x for
any x in E;
(d) For every element x ∈ E there exists a unique element (−x) of the
same space, such that x + (−x) = θ.
(e) The element θ is said to be the null element or zero element of E and
the element −x is called the inverse element of x.
The set E satisfying the axioms (i) and (ii) is called a linear or vector
space. This is said to be a real or complex space depending on whether the
set of multipliers are real or complex.
1.3.1 Examples
(i) Real line 4
The set of all real numbers for which the ordinary additions and
multiplications are taken as linear operations, is a real linear space . 4
(ii) The Euclidean space 4 , unitary space C
n n
, and complex plane
+
Let X be the set of all ordered n-tuples of real numbers. If x =
(ξ1 , ξ2 , . . . , ξn ) and y = (η1 , η2 , . . . , ηn ), we define the operations of addition
and scalar multiplication as x + y = (ξ1 + η1 , ξ2 + η2 , . . . , ξn + ηn ) and
λx = (λξ1 , λξ2 , . . . , λξn ). In the above equations, λ is a real scalar. The
above linear space is called the real n-dimensional space and denoted by n . 4
+
The set of all ordered n-tuples of complex numbers, n , is a linear space
with the operations of additions and scalar multiplication defined as above.
The complex plane + is a linear space with addition and multiplication of
complex numbers taken as the linear operations over 4 +
(or ).
(ii) n4
Consider the real linear space 4 n
where every x ∈ 4 n
is an
ordered n-tuple of real numbers. Let e1 = (1, 0, 0, . . . , 0), e2 =
(0, 1, 0, 0, . . . , 0), . . . , en = (0, 0, . . . , 1). We may note that {ei }, i =
1, 2, . . . , n is a linearly independent set and spans the whole space n . 4
4 4
Hence, {e1 , e2 , . . . , en } forms a basis of n . For n = 1, we get 1 and any
singleton set comprising a non-zero element forms a basis for 1 . 4
+
(iii) n
+
The complex linear space n is a linear space where every x ∈ n is +
an ordered n-tuple of complex numbers and the space is finite dimensional.
+
The set {e1 , e2 , . . . , en }, where ei is the ith vector, is a basis for n .
(v) 4m×n(+m×n )
4m×n is the space of all matrices of order m × n. For i = 1, 2, . . . , n let
Ei×j be the m×n matrix with (i, j)th entry as 1 and all other entries as zero.
Then, {Eij : i = 1, 2, . . . , m; j = 1, 2, . . . , n} is a basis for (4)m×n (C m×n ).
1.3.6 Theorem
Let E be a finite dimensional linear space. Then all the bases of E have
the same number of elements.
Let {e1 , e2 , . . . , en } and {f1 , f2 , . . . , fn , fn+1 } be two different bases in
E. Then, any element fi can be expressed as a linear combination of
10 A First Course in Functional Analysis
n
e1 , e2 , . . . , en ; i.e., fi = aij ej . Since fi , i = 1, 2, . . . , n are linearly
i=1
independent, the matrix [aij ] has rank n. Therefore, we can express fn+1
n
as fn+1 = an+1j ej . Thus the elements f1 , f2 , . . . , fn+1 are not linearly
j=1
independent. Since {f1 , f2 , . . . , fn+1 } forms a basis for the space it must
contain a number of linearly independent elements, say m(≤ n). On the
other hand, since {fi }, i = 1, 2, . . . , n + 1 forms a basis for E, ei can be
expressed as a linear combination of {fj }, j = 1, 2, . . . , n + 1 such that
n ≤ m. Comparing m ≤ n and n ≤ m we conclude that m = n. Hence the
number of elements of any two bases in a finite dimensional space E is the
same.
The above theorem helps us to define the dimension of a finite
dimensional space.
1.3.7 Dimension, examples
The dimension of a finite dimensional linear space E is defined as the
number of elements of any basis of the space and is written as dim E.
For an infinite dimensional space it can be shown that all bases are
equivalent sets.
1.3.8 Theorem
If E is a linear space all the bases have the same cardinal number.
Let S = {ei } and T = {fi } be two bases. Suppose S is an infinite
set and has cardinal number α. Let β be the cardinal number of T . Every
fi ∈ T is a linear combination, with non-zero coefficients, of a finite number
of elements e1 , e2 , . . . , en of S and only a finite number of elements of T are
associated in this way with the same set e1 , e2 , . . . , en or some subset of it.
Since the cardinal number of the set of finite subsets of S is the same as
that of S itself, it follows that β ≤ ℵ0 , β ≤ α. Similarly, we can show that
α ≤ β. Hence, α = β. Thus the common cardinal number of all bases in
an infinite dimensional space E is defined as the dimension of E.
1.3.9 Direct sum
Here we consider the representation of a linear space E as a direct sum
of two or more subspaces. Let E be a linear space and X1 , X2 , . . . , Xn
be n subspaces of E. If x ∈ E has an unique representation of the form
x = x1 + x2 + · · · + xn , xi ∈ Xi , i = 1, 2, . . . , n, then E is said to be the
direct sum of its subspaces X1 , X2 , . . . , Xn . The above representation is
Preliminaries 11
called the decomposition of the element x into the elements of the subspaces
n
X1 , X2 , . . . , Xn . In that case we can write E = X1 ⊕X2 ⊕· · · Xn = ⊕Xi .
i=1
(x + M ) + (y + M )
= (x + y ) + M
y+M
x+y
x+M
y x+m
x M
4
space let d(x, y) be the distance between them in , i.e., d(x, y) = |x − y|.
The concept of distance gives rise to the concept of limit, i.e., {xn } is said
to tend to x as n → ∞ if d(xn , x) → 0 as n → ∞. The concept of continuity
can be introduced through the limiting process. We replace the set of real
4
numbers underlying by an abstract set X of elements (all the attributes of
which are known, but the concrete forms are not spelled out) and introduce
on X a distance function. This will help us in studying different classes of
problems within a single umbrella and drawing some conclusions that are
universally valid for such different sets of elements.
1.4.1 Definition: metric space, metric
A metric space is a pair (X, ρ) where X is a set and ρ is a metric on
X (or a distance function on X) that is a function defined on X × X such
that for all x, y, z ∈ X the following axioms hold:
n
Let, z = (ζ1 , ζ2 , . . . , ζn ). Then, ρ2 (x, z) = (ξi − ζi )2
j=1
n
n
n
= (ξi − ηi )2 + (ηi − ζi )2 + 2 (ξi − ηi )(ηi − ζi )
i=1 i=1 i=1
Now by the Cauchy-Bunyakovsky-Schwartz inequality [see 1.4.3]
n
1/2 n
1/2
n
2 2
(ξi − ηi )(ηi − ζi ) ≤ (ξi − ηi ) (ηi − ζi )
i=1 i=1 i=1
≤ ρ(x, y)ρ(y, z)
Thus, ρ(x, z) ≤ ρ(x, y) + ρ(y, z).
Hence, all the axioms of a metric space are fulfilled. Therefore, n 4
under the metric defined by (1.1) is a metric space and is known as the
n-dimensional Euclidean space. If x, y, z denote three distinct points in 2 4
then the inequality ρ(x, z) ≤ ρ(x, y) + ρ(y, z) implies that the length of any
side of a triangle is always less than the sum of the lengths of the other
two sides of the triangle obtained by joining x, y, z. Hence, axiom 4) of
the set of metric axioms is known as the triangle inequality. n-dimensional
+
unitary space n is the space of all ordered n-tuples of complex numbers
n
with metric defined by ρ(x, y) = (ξi − ηi )2 . When n = 1 this is the
i=1
complex plane + with the usual metric defined by ρ(x, y) = |x − y|.
(iii) Sequence space l∞
Let X be the set of all bounded sequences of complex numbers, i.e.,
every element of X is a complex sequence x = (ξ1 , ξ2 , . . . , ξn ) or x = {ξi }
such that |ξi | ≤ Ci , where Ci for each i is a real number. We define the
metric as ρ(x, y) = sup |ξi − ηi |, where y = {ηi }, N = {1, 2, 3, . . .}, and
i∈N
‘sup’ denotes the least upper bound (l.u.b.). l∞ is called a sequence space
because each element of X (each point in X) is a sequence.
The set of all continuous functions defined on the interval [a, b] with the
above metric is called the space of continuous functions and is defined on
J and denoted by C([a, b]). This is a function space because every point of
C([a, b]) is a function.
n
+ sup |(g(ti ) − h(ti )) − (g(ti−1 ) − h(ti−1 ))|
i=1
≤ V (f − g) + V (g − h) = ρ(f, g) + ρ(g, h)
Thus all the axioms of a metric space are fulfilled.
Axioms 1-3 of a metric space are satisfied. To see that ρ(x, y) also
satisfies axiom 4 of a metric space, we proceed as follows:
t 1
Let f (t) = , t ∈ R. Since f (t) = > 0,
1+t (1 + t)2
f (t) is monotonically increasing.
Hence |a + b| ≤ |a| + |b| ⇒ f (|a + b|) ≤ f (|a|) + f (|b|).
|a + b| |a| + |b| |a| |b|
Thus, ≤ ≤ + .
1 + |a + b| 1 + |a| + |b| 1 + |a| 1 + |b|
16 A First Course in Functional Analysis
Problems
√
1. Show that ρ(x, y) = x − y defines a metric on the set of all real
numbers.
2. Show that the set of all n-tuples of real numbers becomes a metric
space under the metric ρ(x, y) = max{|x1 − y1 |, . . . , |xn − yn |} where
x = {xi }, y = {yi }.
3. Let 4
be the space of real or complex numbers. The distance ρ of
two elements f, g shall be defined as ρ(f, g) = ϕ(|f − g|) where ϕ(x) is
a function defined for x ≥ 0, ϕ(x) is twice continuously differentiable
and strictly monotonic increasing (that is, ϕ (x) > 0), and ϕ(0) = 0.
Then show that ρ(f, g) = 0 if and only if f = g.
4. Let C([B]) be the space of continuous (real or complex) functions f ,
defined on a closed bounded domain B on n . Define ρ(f, g) = ϕ(r) 4
where r = max |f − g|. For ϕ(r) we make the same assumptions as
4 is
B
in example 3. When ϕ (r) < 0, show that the function space
metric, but, when ϕ (r) > 0 the space is no more metric. 4
1.4.3 Theorem (Hölder’s inequality)
1 1
If p > 1 and q is defined by + = 1
p q
n
1/p n
1/q
n
(H1) |xi yi | ≤ |xi | p
|yi |q
for j = 1, 2, . . . n, we get
|xj yj | aj bj
1/p
1/q ≤ p + q , j = 1, 2, . . . , n.
n n
|xj |p |yj |q
j=1 j=1
|xj |p |yj |q
Taking aj =
∞ , bj =
∞ for j = 1, 2, . . . ∞
|xi |p |yi |q
i=1 i=1
we obtain as in above,
|xj yj |p aj bj
∞ 1/p ∞ 1/q ≤ p + q .
|xi | p |yi |q
i=1 i=1
Moreover, (|xi | + |yi |)p = (xi | + |yi |)p−1 |xi | + (|xi | + |yi |)p−1 |yi |.
Summing these identities for i = 1, 2, . . . , n,
n
1/p n
1/q
n
(|xi + yi | )|xi | ≤
p−1
|xi | p
((|xi | + |yi |) )
p−1 q
Similarly we have,
1/p
1/q
n
n
n
(|xi | + |yi |) p−1
|yi | ≤ |yi | p
(|xi | + |yi | ) p
n
1/q
(|xi | + |yi |) p
(1.6)
i=1
n
1/p n
1/p
1/p
n
or, (|xi | + |yi |) p
≤ |xi |
p
+ |yi | p
n
assuming that (|xi | + |yi |)p = 0.
i=1
Assuming |xi + yi |p = 0, the above inequality yields on division
i=1
∞
1/q
by |xi + yi | p
,
i=1
1/p ⎡
1/p
1/p ⎤
∞
∞
∞
|xi + yi |p ≤⎣ |xi |p + |yi |q ⎦ (1.8)
i=1 i=1 i=1
Preliminaries 21
1
Assuming that |x(t) + y(t)|p dt = 0 and dividing both sides of the
0
1 1/q
above inequality by |x(t) + y(t)| dt
p
, we get
0
1 1/p
|x(t) + y(t)| dt p
0
1/p 1/p
1 1
≤ |x(t)| dt
p
+ |y(t)|p dt (1.9)
0 0
Problems
1. Show that the Cauchy-Bunyakovsky-Schwartz inequality implies that
(|ξ1 | + |ξ2 | + · · · + |ξn |)2 ≤ n(|ξ1 |2 + · · · + |ξn |2 ).
2. In the plane of complex numbers show that the points z on the open
unit disk |z| < 1 form a metric space if the metric is defined as
1 1+u z1 − z2
ρ(z1 , z2 ) = log , where u = .
2 1−u 1 − z 1 z2
(n) (n)
1.4.5 The spaces lp , l∞ , lp , p ≥ 1, l∞
(n) (n)
(i) The spaces lp , l∞
Let X be an n-dimensional arithmetic space, i.e., the set of all possible
n-tuples of real numbers and let x = {x1 , x2 , . . . , xn }, y = {y1 , y2 , . . . , yn },
and p ≥ 1.
n
1/p
We define ρp (x, y) = |xi − yi | p
. Let max |xi −yi | = |xk −yk |.
1≤i≤n
i=1
Then,
⎛ ⎞1/p
n p
⎜ xi − yi ⎟
ρp (x, y) = |xk − yk | ⎝1 +
xk − y k ⎠ .
i=1
i=k
axiom 4 of a metric space is satisfied. Hence the set X with the metric
(n)
ρp (x, y) is a metric space and is called l∞ .
Preliminaries 23
limits are in the sense of uniform convergence. However, F ([0, 1]) admits
of only pointwise convergence. This means int M = Φ, i.e., M is nowhere
dense, that is therefore a first category set of functions [sec 1.4.9.]. Thus,
M is the set of real functions and their limits are in the sense of pointwise
convergence. Therefore, M is a set of the second category [sec. 1.4.9] and
the pointwise convergence is non-metrizable.
Problems
1. Find a sequence which converges to 0, but is not in any space lp where
1 ≤ p < ∞.
|x − y|
2. Show that the real line with ρ(x, y) = is a metric space.
1 + |x − y|
3. If (X, ρ) is any metric space, show that another metric of X is defined
by
ρ(x, y)
ρ(x, y) = .
1 + ρ(x, y)
Note 1.4.2. In what follows we show how different metrics yield different
4
types of open balls. Let X = 2 be the Euclidean space. Then the unit
open ball B(0, 1) is given in Figure 1.2(a). If the l∞ norm is used the unit
open ball B(0, 1) is the unit square as given in Figure 1.2(b). If the l1 norm
is used, the unit open ball B(0, 1) becomes the ‘diamond shaped’ region
shown in Figure 1.2(c). If we select p > 2, B(0, 1) becomes a figure with
curved sides, shown in Figure 1.2(d). The unit ball in C[0, 1] is given in
26 A First Course in Functional Analysis
Figure 1.2(e).
x2 x2 x2
x1 x1 x1
x1 f0 + r
2r
f0
f0−r
B (0,1)⊂(R 2, ρl3) x1
1
(d) (e)
Fig. 1.2
Note 1.4.3. It may be noted that the closed sets of a metric space have
the same basic properties as the closed numerical point sets, namely:
(i) M ∪ N = M ∪ N ;
(ii) M ⊂ M ;
(iii) M = M = M ;
(iv) The closure of an empty set is empty.
1.4.9 Theorem
In any metric space X, the empty set Φ and the full space X are open.
To show that Φ is open, we must show that each point in Φ is the centre
of an open ball contained in Φ; but since there are no points in Φ, this
requirement is automatically satisfied. X is clearly open since every open
ball centered on each of its points is contained in X. This is because X is
the entire space.
Note 1.4.4. It may be noted that an open ball B(0, r) on the real line is
the bounded open interval ] − r, r[ with its centre as the origin and a total
length of 2r. We may note that [0,1[ on the real line is not an open set since
the interval being closed on the left, every bounded open interval with the
origin as centre or in other words every open ball B(0, r) contains points
of 4not belonging to [0, 1[. On the other hand, if we consider X = [0, 1[
as a space itself, then the set X is open. There is no inconsistency in the
above statement, if we note that when X = [0, 1[ is considered as a space,
Preliminaries 27
there are no points of the space outside [0,1[. However, when X = [0, 1[ is
4 4
considered as a subspace of , there are points of outside X. One should
take note of the fact that whether or not a set is open is relevant only with
respect to a specific metric space containing it, but never on its own.
1.4.10 Theorem
In a metric space X each open ball is an open set.
Let B(x0 , r) be a given ball in a metric space X. Let x be any point in
B(x0 , r). Now ρ(x0 , x) < r. Let r1 = r−ρ(x0 , x). Hence B(x, r1 ) is an open
ball with centre x and radius r1 . We want to show that B(x, r1 ) ⊆ B(x0 , r).
For if y ∈ B(x, r1 ), then
ρ(y, x0 ≤ ρ(y, x) + ρ(x, x0 ) < r1 + ρ(x, x0 ) = r, i.e., y ∈ B(x0 , r).
Thus B(x, r1 ) is an open ball contained in B(x0 , r). Since x is arbitrary, it
follows that B(x0 , r) is an open set. In what follows we state some results
that will be used later on:
Let X be a metric space.
(i) A subset G of X is open ⇔ it is a union of open balls.
(ii) (a) every union of open sets in X is open and (b) any finite
intersection of open sets in X is open.
Note 1.4.5. The two properties mentioned in (ii) are vital properties
of a metric space and these properties are established by using only the
‘openness’ of a set in a metric space. No use of distance or metric is required
in the proof of the above theorem. These properties are germane to the
development of ‘topology’ and ‘topological spaces’. We discuss them in the
next section.
We will next mention some properties of closed sets in a metric space.
We should recall that a subset K of X is said to be closed if its complement
K C = X − K is open.
(i) The null set Φ and the entire set X in a metric space are closed.
(ii) In a metric space a closed ball is a closed set.
(iii) In a metric space, (a) the intersection of closed sets is closed and
(b) the union of a finite number of closed sets is a closed set.
Note 1.4.10. The converse of the theorem is not true for an arbitrary
metric space, since there exist metric spaces that contain a Cauchy sequence
but have no element that will be the limit.
Preliminaries 29
1.4.15 Examples
(i) The space of rational numbers
Let X be the set of rational numbers, in which the metric is taken
as ρ(r1 , r2 ) = |r1 − r2 |. Thus, X is a metric space. Let us take
r1 = 1, r2 = 12 , · · · , rn = n1 . {rn } is a Cauchy sequence and rn → 0 as
! "n
n → ∞. On the other hand let us take rn = 1 + n1 where n is an
! "
1 n
integer. {rn } is a Cauchy sequence. However, lim 1 + n = e, which is
n→∞
not a rational number.
(p) (q)
Squaring, we have for p, q > n0 (
), i = 1, 2, . . . , n, |ξi − ξi |2 <
2 ⇒
(p) (q)
|ξi − ξi | <
. This shows that for each fixed i, (1 ≤ i ≤ n),
(i) (i) (i)
the sequence {ξ1 , ξ2 , . . . , ξn } is a Cauchy sequence of real numbers.
Therefore, ξi
(m)
= ξi ∈4 as m → ∞. Let us denote by x the vector,
4
x = (ξ1 , ξ2 , . . . , ξn ). Clearly, x ∈ n . It follows from (1.1) that ρ(xm , x) ≤
for m ≥ n0 (
). This shows that x is the limit of {xm } and this proves
completeness because {xm } is an arbitrary Cauchy sequence. Completeness
+
of n can be proven in a similar fashion.
fixed t = t0 ∈ J = [a, b], |xn (t0 ) − xm (t0 )| <
for m, n > n0 (
). Thus,
{xn (t0 )} is a convergent sequence of real numbers. Since 4
is complete,
4
{xn (t0 )} → x(t0 ) ∈ . In this way we can associate with each t ∈ J a
unique real number x(t) as limit of the sequence {xn (t)}. This defines a
(pointwise) function x on J and thus x(t) ∈ C([a, b]). Thus, it follows from
(6.2), making n → ∞, max |xm (t) − x(t)| <
for m ≥ n0 (
) for every t ∈ J.
Therefore, {xm (t)} converges uniformly to x(t) on J. Since xm (t)’s are
continuous functions of t, and the convergence is uniform, the limit x(t) is
continuous on J. Hence, x(t) ∈ C([a, b]), i.e., C([a, b]), is complete.
Note 1.4.11. We would call C([a, b]) as real C([a, b]) if each member of
C([a, b]) is real-valued. On the other hand, if each member of C([a, b]) is
+
complex-valued then we call the space as complex ([a, b]).
By arguing analogously as above we can show that complex ([a, b]) is +
complete. We next consider the set X of all continuous real-valued functions
on J = [a, b]. Let us define the metric ρ(x(t), y(t)) for x(t), y(t) ∈ X as
b
ρ(x, y) = |x(t) − y(t)|dt.
a
We can easily see that the set X with xm (t ), xn (t )
the metric defined above is a metric space
S[a, b] = (X, ρ). We next show that O
S[a, b] is not complete. Let us construct
a {xn } as follows: If a < c < b and for
every n so large that a < c − n1 1
We define ⎧ A B C
⎪ 1 0 t
⎪
⎪ 0 if a ≤ t ≤ c − 1
c −—
1
c −—
⎪
⎨ n m n
xn (t) = 1
⎪
⎪ nt − nc + 1 if c − ≤ t ≤ c Fig. 1.3
⎪
⎪ n
⎩
1 if c ≤ t ≤ b
For n > m
b
1 1 1 1 1 1
|xn (t) − xm (t)|dt = ΔAOB = · 1 · c − − c + < +
a 2 n m 2 n m
4. Let X be a metric space and B(x, r) the open ball in X with centre x
and radius r. Let A be a subset of X with diameter less than r that
intersects B(x, r). Prove that A ⊆ B(x, 2r).
5. Show that the closure B(x0 , r) of an open ball B(x0 , r) in a metric
space can differ from the closed ball B(x0 , r).
6. Describe the interior of each of the following subsets of the real line:
the set of all integers; the set of rationals; the set of all irrationals;
]0, 1]; [0, 1]; and [0, 1[∪{1, 2}.
7. Give an example of an infinite class of closed sets, the union of which
is not closed. Give an example of a set that (a) is both open and
closed, (b) is neither open nor closed, (c) contains a point that is not
a limit point of the set, and (d) contains no points that are not limit
points of the set.
8. Describe the closure of each of the following subsets of the real line;
the integers; the rationals; ]0, +∞[; ] − 1, 0[∪]0, 1[.
9. Show that the set of all real numbers constitutes an incomplete metric
space if we choose ρ(x, y) = | arctan x − arctan y|.
10. Show that the set of continuous real-valued functions on J = [0, 1]
do not constitute a complete metric space with the metric ρ(x, y) =
1
|x(t) − y(t)|dt.
0
11. Let X be the metric space of all real sequences x = {ξi } each
of which
has only finitely many nonzero terms, and ρ(x, y) = |ξi − ηi |,
(n) (n) (n)
where y = {ηi }. Show that {xn } with xn = {ξj }, ξj = j −2 for
(n)
j = 1, 2, . . . , n, and ξj = 0 for j > n is a Cauchy sequence but does
not converge.
12. Show that {xn } is a Cauchy sequence if and only if ρ(xn+k , xn )
converges to zero uniformly in k.
13. Prove that the sequence 0.1, 0.101, 0.101001, 0.1010010001,. . . is a
Cauchy sequence of rational numbers that does not converge in the
space of rational numbers.
' (
14. In the space l2 , let A = x = (x1 , x2 , . . .) : |xn | ≤ n1 , n = 1, 2, . . . .
Prove that A is closed.
There is an N = N (
h) such that k > N implies that sup |fk (t)−f (t)| ≤
t
h
. Since f is continuous, there is an M > N such that for kj > M , we
4
h
h
have, |f (ξ+h)−f (ξkj +h)| < and |f (ξkj −f (ξ)| < . Since lim ξkj = ξ,
4 4 j→∞
f (ξ + h) − f (ξ) fkj (ξkj + h) − fkj (ξkj )
if kj > M , we have <
+
≤
h h
f (|ξ + h) − f (ξ)
n +
. It follows that ≤ n for all h > 0. Thus f ∈ Kn
h
and Kn is closed.
For (ii) Let us suppose that g ∈ Γ. Let
> 0 and let us partition [0,
1] into k equal intervals such that if x, x are in the same interval of the
partitioning, |g(x) − g(x )| <
/2 holds. Let us consider th
! i−1 "the i !subinterval
"
i−1
≤ x ≤ k and consider the rectangle with sides g k and g ki . For all
i
k ! " ! "
points!within! the rectangle the ordinates satisfy g i−1
"" k − 2 ≤ !y ≤ g !k + ""2.
Thus ki , g ki is on the right-hand side of the rectangle and i−1 k , g i−1
k
is on the left-hand side of the rectangle. By joining these two points by a
polygonal graph that remains within the rectangle and the line segments of
which have slopes exceeding n in absolute value, we thus obtain a continuous
function that is within
of g and as because its slope exceeds n, it belongs
to Γ ∼ Kn . Thus Γ − Kn is dense in Γ. Combining (i) and (ii) we can say
that Kn and hence K is nowhere dense in Γ.
x2
ε/2
ε/2 { { g (x)
x1
i −1 i
—— —
k k
Fig. 1.4
Preliminaries 35
+ · · · + ρ(xn+p−1 , xn+p )
≤ (α + α
n n+1
+ ··· + α n+p−1
)ρ(x0 , A(x0 ))
α −α
n n+p
= ρ(x0 , A(x0 )).
1−α
n
Since, by hypothesis, ρ(xn , xn+p ) ≤ 1−α α
ρ(x0 , A(x0 )), therefore
ρ(xn , xn+p ) → 0 as n → ∞, p > 0. Thus, (xn ) is a Cauchy sequence.
Since the space is complete, there is an element x∗ ∈ X, the limit of the
sequence, x∗ = lim xn .
n→∞
For n sufficiently large, we can write, ρ(x∗ , xn ) <
/2, ρ(x∗ , xn−1 ) <
/2α, for any given
. Hence ρ(x∗ , A(x∗ )) <
. Since
> 0 is arbitrary,
ρ(x∗ , A(x∗ )) = 0, i.e., A(x∗ ) = x∗ .
Let us assume that there exists two elements, x∗ ∈ X, y ∗ ∈ Y, x∗ = y ∗
satisfying A(x∗ ) = x∗ and A(y ∗ ) = y. Then, ρ(x∗ , y ∗ ) = ρ(A(x∗ ), A(y ∗ )) ≤
αρ(x∗ , y ∗ ). Since x∗ = y ∗ , and α < 1, the above inequality is
impossible unless ρ(x∗ , y ∗ ) = 0, i.e, x∗ = y ∗ . Making p → ∞, in
n
−αn+p
the inequality ρ(xn , xn+p ) ≤ α 1−α ρ(x0 , A(x0 )), we obtain ρ(xn , x∗ ) ≤
αn
1−α ρ(x0 , A(x0 )).
Note 1.4.12. Given an equation F (x) = 0, where F : n → n , we can 4 4
write the equation F (x) = 0 in the form x = x − F (x). Denoting x − F (x)
by A(x), we can see that the problem of finding the solution of F (x) = 0 is
equivalent to finding the fixed point of A(x) and vice versa.
1.4.27 Applications
(i) Solution of a system of linear equations by the iterative
method
Let us consider the real n-dimensional space. If x = (ξ1 , ξ2 , . . . , ξn ) and
y = (η1 , η2 , . . . , ηn ), let us define the metric as ρ(x, y) = max |ξi − ηi |. Let
i
us consider y = Ax, where A is an n × n matrix, i.e., A = (aij ). The
n
system of linear equations is given by ηi = aij ξj , i = 1, 2, . . . , n. Then
j=1
(1) (2)
(1) (2)
,
ρ(y1 2y ) = ρ(Aξ , Aξ ) yields max |η −η | = max a (ξ − ξ ) ≤
i
1 2 i i ij j j
i
j
n
max |aij |ρ(x1 , x2 ). Now if it is assumed that |aij | < 1, for all i, then
i
j=1 j
38 A First Course in Functional Analysis
1.4.28 Theorem
Then the integral equation x(t) = f (t) + λ k(t, s)x(s)ds has a unique
a
solution x(t) ∈ L2 ([a, b]) for every sufficiently small value of the parameter
λ.
b
Proof: Consider the operator Ax(t) = f (t)+λ k(t, s)x(s)ds. Let x(t) ∈
b a
L2 ([a, b]), i.e., x2 (t)dt < ∞. We first show that for x(t) ∈ L2 ([a, b]),
a
Ax ∈ L2 ([a, b]).
b b b b
2 2
(Ax) dt = f (t)dt + 2λ f (t) k(t, s)x(s)ds dt
a a a a
2
b b
2
+λ k(t, s)x(s)ds dt
a a
Using Fubini’s theorem th. 10.5 and the square integrability of k(t, s) we
can show that
b b b b
f (t) k(t, s)x(s)ds dt = k(t, s)x(s)f (t)dt ds
a a a a
1/2
1/2
1/2
b b b b
2 2 2
≤ k (t, s)dt ds f (t)dt · x (s)ds
a a a a
< +∞
2
b b
Similarly we have k(t, s)x(s)ds dt < ∞. Thus, A(x) ∈
a a
L2 ([a, b]). Therefore, A : L2 ([a, b]) → L2 ([a, b]). Using the metric in
L2 ([a, b]), i.e., given x(t), y(y) ∈ L2 ([a, b]),
1/2
b
ρ(Ax, Ay) = |Ax − Ay|2 dt
a
Preliminaries 39
⎡
2 ⎤1/2
b b
= |λ| ⎣ k(t, s)(x(s) − y(s))ds dt⎦
a a
2
b b b
≤ |λ| |k(t, s)|2 dt ds |x(s) − y(s)|2 ds 1/2
a a a
1/2
b b
2
= |λ| |k(t, s)| dt ds ρ(x, y) < αρ(x, y)
a
1/2
b b
2
where α = |λ| |k(t, s)| dt ds and α < 1 if |λ| <
aa
−1/2
b b
|k(t, s)2 |dt ds . Thus the contraction mapping principle
a a
holds, proving the existence and uniqueness of the solution of the given
integral equation for values of λ satisfying the above inequality.
2ξ1 + ξ2 + ξ3 = 4
ξ1 + 2ξ2 + ξ3 = 4
ξ1 + ξ2 + 2ξ3 = 4
12. Show that x = 3x2/3 , x(0) = 0 has infinitely many solutions, x, given
by x(t) = 0 if t < c and x(t) = (t − c)3 if t ≥ c, where c > 0 is
any constant. Does 3x2/3 on the right-hand side satisfy a Lipschitz
condition?
13. Pseudometric: A finite pseudometric on a set X is a function
ρ : X×X → 4
satisfying for all x ∈ X conditions (1), (3) and
(4) of Section 1.4.1 and 2 (i.e. ρ(x, x) = 0, for all x ∈ X).
Preliminaries 43
> 0 being any preassigned number. On the other hand, there exists
another polynomial p0 (t) with rational coefficients, s.t.,
Hence ρ(x, p0 ) = max |x(t) − p0 (t)| <
. Hence C([0, 1]) is separable.
t
Note 1.4.14. The Weierstrass theorem for C([0, 1]) says in effect that all
real linear combinations of functions 1, x, x2 , . . . , xn are dense in [0, 1].
The Separability of the Space lp (1 < p < ∞)
Let E0 be the set of all elements x of the form (r1 , r2 , . . . , rn ) where ri
are rational numbers and n is an arbitrary natural number. E0 is countable.
We would like to show that E0 is everywhere dense in lp . Let us take an
element x = {xi } ∈ lp and let an arbitrary
> 0 be given. We find a
∞
p
natural number n0 such that for n > n0 |ξk |p < . Next, take an
2
k=n+1
∞
p
element x0 = (r1 , r2 , . . . , rn,0,0 . . .) such that |ξk − rk |p < . Then,
2
k=1
∞
∞
p
p
p
[ρ(x, x)] = |ξk − rk | +
p
|ξk | <
p
+ =
p where ρ(x, x0 ) <
.
2 2
k=1 k=n+1
1.5.2 Example
Let X = (α1 , α2 , α3 ). Consider 1 = {φ, X, {α1 }, {α1 , α2 }}, 2 =
{φ, X, {α1 }, {α2 }}, 3 = {φ, x, {φ1 }, {α2 }, {α3 }, {α1 , α2 }, {α2 , α3 }} and
4 = {φ, X}.
Here, 1 , 3 , and 4 are topologies, but 2 is not a topology due to
the fact that {α1 } ∪ {α2 } = {α1 , α2 } ∈ 2 .
1.5.3 Definition: indiscrete topology, discrete topology
An indiscrete topology denoted by ‘J’, has only two members, Φ and
X. The topological space (X, J) is called an indiscrete topological space.
Another trivial topology for a non-empty set X is the discrete topology
denoted by ‘D’. The discrete topology for X consists of all subsets of X.
A topological space (X, D) is called a discrete topological space.
1.5.4 Example
Let X = R. Consider the topology ‘S’ where Φ ∈ S. If G ⊆ R
and G = Φ, then G ∈ S if for each p ∈ G there is a set H of the form
{x ∈ R : a ≤ x < b}, a < b, such that p ∈ H and H ⊆ G. The set
H = {x ∈ R : a ≤ x < b} is called a right-half open interval. Thus a
nonempty set G is S-open if for each p ∈ G, there is a right-half open
interval H such that p ∈ H ⊆ G. The topology defined above is called a
limit topology.
1.5.5 Definition: usual topology, upper limit topology, lower
limit topology
Let X = R {real}. Let us consider a topology = {Φ, R, {]a, b[}, a < b
and all unions of open intervals} on X = R. This type of topology is called
the usual topology.
Let X = R be the non-empty set and = {φ, R, {]a, b]}, a < b and
union of left-open right closed intervals}. This type of topology is called
the upper limit topology.
Let X = R be the non-empty set and = {Φ, X = R, {[a, b[}, a < b
and union of left-closed right-open intervals}. Then this type of topology
is called lower limit topology.
1.5.6 Examples
(i) (Finite Complement Topology)
Let us consider an infinite set X and let = {Φ, X, A ⊂ X|AC be a
finite subset of X}. Then we see that is a topology and we call it a finite
complement topology.
(ii) (Countable complement topology)
Let X be a non-enumerable set and = {Φ, X, A ⊂ X|AC be a
countable complement}. Then is a topology and will be known as a
countable complement topology.
(iii) In the usual topology in the real line, a single point set is closed.
Preliminaries 47
Closure of a set
If A is a subset of a topological space, then the closure of A (denoted
by A) is the intersection of all closed supersets of A. It is easy to see that
the closure of A is a closed superset of A that is contained in every closed
superset of A and that A is closed ⇔ A = A.
A subset of a topological space X is said to dense or everywhere dense
if A = X.
1.5.9 Definition: neighbourhood
Let (X, ) be a topological space and x ∈ X be an arbitrary point. A
subset Nx ⊆ X containing the point x ∈ X is said to be a neighbourhood
of x if ∃ a -open set Gx such that x ∈ Gx ⊆ Nx .
Clearly, every open set Gx containing x is a neighbourhood of x.
1.5.10 Examples
4
(i)]x −
1 , x +
2 [ is an open neighbourhood in ( , U ).
4
(ii)]x, x +
] is an open neighbourhood in ( , UL ), where UL is a topology
defined above . 4
4
(iii)[x −
, x[ is an open neighbourhood in ( , U4 ).
1.5.20 Example
4 4
( , U ) is first countable. ( , UL ) is second countable. It is to be noted
that a second countable space is also first countable.
Problems
1. For each p ∈ 4find a collection Bp such that Bp is a base for the
D-neighbourhood system of p.
2. Let X = {a, b, c} and let = {X, Φ, X, {a}, {b, c}}. Show that is a
topology for X.
3. For each p ∈ X find a collection Bp of basic -neighbourhoods of p.
4. Prove that open rectangles in the Euclidean plane form an open base.
Definition: limit point, contact point, isolated point, derived set, closure
Let (X, ) be a topological space and let A be a subset of X. The
point x ∈ X is said to be a limit point of A if every -neighbourhood
of x contains at least one point of A other than x. That is, x is a limit
point of A if and only if Nx a -neighbourhood of x satisfies the condition
Nx ∩ (A − {x}) = Φ.
If ∀ neighbourhoods Nx of x, s.t. Nx ∩ A = Φ, then x is called a contact
point of A. D(A) = {x : x is a limit point of A} is called the derived set of
A.
A ∪ D(A) is called the closure of A denoted by A.
Problems
4
Let X = and A =]0, 1[. Then find D(A) for the following cases:
(i) = U , the usual topology on 4
(ii) = UL , the lower limit topology on 4
4
(iii) = U4 , the upper limit topology on .
1.5.22 Theorem
Let (X, ) be a topological space. Let A be a subset of X and D(A) the
set of all limit points of A. Then A ∪ D(A) is -closed.
1.5.23 Definition: -closure
Let (X, ) be the topological space and A be a subset of X. The -
closure of A denoted by A is the smallest -closed subset of X that contains
A.
Preliminaries 51
1.5.24 Theorem
Let (X, ) be a topological space and let A be a subset of X. Then
A = A ∪ D(A) where D(A) is the set of limit points of A. It follows from
the previous theorem that A ∪ D(A) is a closed set.
1.5.25 Definition: interior, exterior, boundary
Let (X, ) be a topological space and let A be a subset of X. A point x
is a -interior point of A if A is a -neighbourhood of x. The -interior
of A denoted by Int A is the set of all interior points of A.
x ∈ X is said to be an exterior point of A if x is an interior point of
AC = X ∼ A.
A point in X is called a boundary point of A if each -neighbourhood
of x contains points both of A and of AC . The -boundary of A is the set
of all boundary points of A.
1.5.26 Example
4
Consider the space ( , U ). The point
1
2
is an interior point of [0, 1],
but neither 0 nor 1 is an interior point of [0, 1]. The U -interior of [0, 1] is
4
]0, 1[. In the UL -topology for , ‘0’ is an interior point of [0, 1] but 1 is not.
The UL -interior of [0, 1] is [0, 1[.
1.5.27 Definition: separable space
Let (X, ) be a topological space. If ∃ a denumerable (enumerable)
subset A of X, A ⊆ X such that A = X, then X is called a separable
space. Or, in other words, a topological space is said to be separable if
it contains a denumerable everywhere dense subset.
1.5.28 Example
4
Let X = and A be the set of all intervals. Now let D(A) be the derived
4
set of A. It is clear that D(A) = . Hence A = A ∪ D(A) = A ∪ = X. 4
4
Hence A is everywhere dense in X = . Similarly, the set 3
of rational
4 4
points, which is also a subset of , is everywhere dense in . Again, since
3 is countable the topological space 4
is separable.
1.6.6 Theorem
Let I1 and I2 be any two open intervals. Then (I1 , U1 ) and (I2 , U2 ) are
homeomorphic, and a homeomorphism exists between the spaces (I1 , UI1 )
and (I2 , UI2 ).
Preliminaries 53
x2
Iu1 h
x1
Iu2
Fig. 1.5
Note 1.6.2. The outcome of the Heine-Borel theorem on the real line is
taken as a definition of compactness in a topological space. The Heine-
Borel theorem reads as follows: If X is a closed and bounded subset of the
4 4
real line , then any class of open subsets of , the union of which contains
X, has a finite subclass whose union also contains X.
1.6.9 Theorem
4
The space ( , U ) is not compact. Therefore, no open interval is compact
w.r.t. the U -topology.
1.6.10 Definition: finite intersection property
Let (X, ζ) be a topological space and {Fλ |λ ∈ Λ} be a class of subsets
such that the intersection of finite number of elements of {Fλ |λ ∈ Λ} is non-
-
n
void, i.e., Fλk = Φ irrespective of whatever manner any finite number
k=1
of λk s {λk |k = 1, 2, . . . , n} is chosen from Λ. Then, F = {Fλ |λ ∈ Λ} is said
to have the Finite Intersection Property (FIP).
1.6.11 Theorem
The topological space (X, ζ) is compact if every class of closed subsets
{Fλ |λ ∈ Λ} possessing the finite intersection property has non-void
54 A First Course in Functional Analysis
-n
intersection. In other words, if all Fλ ’s are closed and Fλk = Φ for a
- k=1
finite subcollection {λ1 , λ2 , . . . , λn }, then Fλ = Φ. The converse result
λ∈Λ
is also true, i.e., if every class of closed subsets of (X, ζ) having the FIP
has non-void intersection, then (X, ζ) is compact.
1.6.12 Theorem
A continuous image of a compact space is compact.
1.6.13 Definition: compactness in metric spaces
A metric space being a topological space under the metric topology,
the definition of compactness as given in definition 1.6.5 is valid in metric
spaces. However, the concept of compactness in a metric space can also be
introduced in terms of sequences and can be related to completeness.
1.6.14 Definition: a compact metric space
A metric space (X, ρ) is said to be compact if every infinite subset of X
has at least one limit point.
⎛
For 1 ≤ p < ∞ and x, y
⎞ p1
∈ 4n(+n), consider ρp (x, y) =
n
⎝ |xj − yj |p ⎠ .
j=1
4 +
(a) (Heine-Borel) A subset of n ( n ) is compact if and only if it is
closed and bounded.
4 +
(b) (Bolzano-Weierstrass) Every bounded sequence in n ( n ) has a
convergent subsequence.
4 +
Proof: Since a bounded subset of n ( n ) is totally bounded, part (a)
4 +
follows from theorem 1.6.18. Since the closure of a bounded set of n ( n )
is complete and totally bounded, part (b) follows from theorem 1.6.18.
1.6.20 Theorem
In a metric space (X, ρ) the following statements are true: X is totally
bounded ⇒ X is separable.
X is compact ⇒ X is separable.
X is separable ⇒ X is second countable.
1.6.21 Theorem
If f is a real-valued continuous function defined on a metric space (X, ρ),
then for any compact set A ⊆ X the values Sup[f (x), x ∈ A], Inf [f (x), x ∈
A] are finite and are attained by f at some points of A.
indicates that f (x) > 0 for every x = x(t) continuous curve that joins the
points (0,0) and (1,1). The fallacy is that the set of curves considered is
not compact even if the set is closed and bounded in C ([0,1]).
1.6.22 Definition: uniformly bounded, equicontinuous
Let (X, ρ) be a complete metric space. The space of continuous real-
valued functions on X with the metric, ρ(x, y) = max[|f (x) − g(x)|x ∈ X]
is a complete metric space, which we denote by C([X]). A collection F of
functions on a set X is said to be uniformly bounded if there is an M > 0
such that |f (x)| ≤ M ∀ x ∈ X and all f ∈ F. For subsets of C([X]),
uniform boundedness agrees with boundedness in a metric space, i.e., a set
F is uniformly bounded if and only if it is contained in a ball.
A collection F of functions defined on a metric space X is called
equicontinuous if for each
> 0, there is a δ > 0 such that ρ(x, x ) <
δ ⇒ |f (x) − f (x )| <
for all x, x ∈ X and for f ∈ F . It may be noted
that the functions belonging to the equicontinuous collection are uniformly
continuous.
1.6.23 Theorem (Arzela-Ascoli)
If (X, ρ) is a compact metric space, a subset K ⊆ C(X) is relatively
compact if and only if it is uniformly bounded and equicontinuous.
1.6.24 Theorem
If f is continuous on an open set D, then for every (x0 , y0 ) ∈ D
dy
the differential equation = f (x, y) has a local solution passing through
dx
(x0 , y0 ).
Problems
1. Show that if (X, ζ) is a topological space such that X has only a finite
number of points, then (X, ζ) is compact.
2. Which of the following subspaces of (R, U ) are compact: (i) J, (ii)
[0, 1], (iii) [0, 1] ∪ [2, 3], (iv) the set of all rational numbers or (v) [2, 3[.
3. Show that a continuous real or complex function defined on a compact
space is bounded. More generally, show that a continuous real or
complex function mapping compact space into any metric space is
bounded.
4. Show that if D is an open connected set and a differential equation
dy
= f (x, y) is such that its solutions form a simple covering of D,
dx
then f is the limit of a sequence of continuous functions.
5. Prove the Heine-Borel theorem: A subspace (Y, UY ) of (R, U ) is
compact if and only if Y is bounded and closed.
1
6. Show that the subset I2 of points {xn } such that |xn | ≤ , n = 1, 2, . . .
n
is compact.
Preliminaries 57
7. Show that the unit ball in C([0, 1]) of points [x : max |x(t)| ≤ 1, t ∈
[0, 1]] is not compact.
8. Show that X is compact if and only if for any collection F of closed
-
n
sets with the property that Fi = Φ for any finite collection in F,
i=1
it follows that ∩{Fλ , Fλ ∈ F} = Φ.
CHAPTER 2
NORMED LINEAR
SPACES
2.1.1 Definition
Let E be a linear space over 4 (or +).
To every element x of the linear space E, let there be assigned a unique
real number called the norm of this element and denoted by ||x||, satisfying
the following properties (axioms of a normed linear space):
58
Normed Linear Spaces 59
Remark 2.1:
x ||x||
Fig. 2.1(a) Fig. 2.1(b)
2.1.2 Examples
1. The n-dimensional Euclidean space 4 n
and the unitary
+
space n are normed linear spaces.
Define the sum and product of the elements by a scalar as in Sec. 1.4.2,
4
Ex. 2. The norm of x ∈ n is defined by
n
1/2
2
||x|| = |ξi | .
i=1
This norm satisfies all the axioms of 2.1.1. Hence 4n and +n are
normed linear spaces.
2. Space C([0, 1]).
We define the addition of functions and multiplication by a scalar
in the usual way.
We set ||x|| = max |x(t)|. It is clear that the axioms of a normed
t∈[0,1]
linear space are satisfied.
60 A First Course in Functional Analysis
It is clear that all the axioms of 2.1.1 are fulfilled. Therefore C k ([0, 1])
is a normed linear space.
Normed Linear Spaces 61
Defining the sum of the elements and the product of elements by a scalar
as in 1.3 and taking the norm as
1/p
p
||x|| = |ξi | p
i=1
∞
1/p ∞
1/p ∞
1/p
(x)
|ξi | p
≤ |ξin |p + |ξi − ξi | p
C([0, 1]) is called the function space. Consider the linear space of all
scalar valued (real or complex) continuous functions defined in C([0, 1]).
Let {xn (t)} ⊂ C([0, 1]) and {xn (t)} be a Cauchy sequence in C([0, 1]).
C([0, 1]) being a normed linear space [see 2.1],
4. l∞ is a Banach space
4+
This shows that for each i, {ξim } is a Cauchy sequence in ( ). Since
4+ 4+
( ) is complete, {ξim } converges in ( ). Let ξim → ξi as m → ∞, and
let x = (ξ1 , ξ2 , . . . , ξn , . . . , . . .).
Let m → ∞, then
(m)
|ξi − ξi | <
∀ m ≥ N (i = 1, 2, . . .) (2.3)
Since the RHS is true for each i and is independent for each i, it follows
that {ξi } is a bounded sequence of numbers and thus x ∈ l∞ . Furthermore,
it follows from (2.3),
||xm − x||∞ <
∀ m > N.
In note 1.4.10, we have shown that the metric space (X, ρ) is not
complete, i.e., given a Cauchy sequence {xm } in (X, ρ), xm does not
64 A First Course in Functional Analysis
⎧ . /
⎪
⎨ 0 ! if t ∈ 0, 12
" . /
xm (t) = m t − 12 if t ∈ 12 , am
⎪
⎩
1 if t ∈ [am , 1]
1 1
where am = + .
2 m
Let us take a = 0 and b = 1 and n > m.
1
Hence, ||xn − xm ||2 = [xn (t) − xm (t)]2 dt
0
1 1
2+n
1 1
2+m
= |xm (t) − xn (t)|2 dt + |xn (t) − xm (t)|dt
1 1 1
2 2+n
1 1 1
= ABC = − < [see figure 2.1(c)].
m n m
The Cauchy sequence does not converge to a point in (X, || · ||). For
every x ∈ X,
1
||xn − x|| = |xn (t) − x(t)|2 dt
0
1/2 am 1
= |x(t)|2 dt + |xn (t) − x(t)|2 dt + |1 − x(t)|2 dt.
0 1/2 am
With the help of Lebesgue integrals the space Lp ([0, 1]) can also be
obtained in a direct way by the use of Lebesgue integral and Lebesgue
measurable functions x on [0, 1] such that the Lebesgue integral of |x|p on
[0, 1] exists and is finite. The elements of Lp ([0, 1]) are equivalent classes of
those functions, where x is equivalent to y if the Lebesgue integral of |x−y|p
over [0, 1] is zero. We discuss these (Lebesgue measures) in Chapter Ten.
Until then the development will take place without the use of measure
theory.
1
— 1
m —
m
1
—
n
C
B
1 1
xn
xm
A
0 0
1 am 1 1 1
— —
2 2
t
Fig. 2.1(c) Fig. 2.1(d)
7. Space s
∞
1 |ξi − ηi |
ρ(x, y) = ,
i=1
2i 1 + |ξi − ηi |
66 A First Course in Functional Analysis
Proof: We have
Problems
4n .
i=1
on
5. Let E be the linear space of all ordered triplets x = {ξ1 , ξ2 , ξ3 }, y =
{η1 , η2 , η3 } of real numbers. Show that the norms on E are defined
by,
||x||1 = |ξ1 | + |ξ2 | + |ξ3 |, ||x||2 = {ξ12 + ξ22 + ξ32 }1/2 ,
||x||∞ = max{|ξ1 |, |ξ2 |, |ξ3 |}
6. Show that the norm is continuous on the metric space associated with
a normed linear space.
7. In case 0 < p < 1, show with the help of an example that ||·||p does not
n
p1
(n)
define a norm on lp unless n = 1 where ||x||p(n) = |ξi |p , x=
i=1
{ξi }.
8. Show that each of the following defines a norm on 42 .
Normed Linear Spaces 67
|x1 | |x2 | x21 x22
(i) |x||1 = + , (ii) |x||2 = + 2 ,
a b a2 b
,
|x1 | |x2 |
(iii) ||x||∞ = max +
a b
where a and b are two fixed positive real numbers and x = (x1 , x2 ) ∈
4 2
. Draw a closed unit sphere (||x|| = 1) corresponding to each of
these norms.
9. Let || · || be a norm on a linear space E. If x + y ∈ E and
||x + y|| = ||x|| + ||y||, then show that ||sx + ty|| = s||x|| + t||y||,
for all s ≥ 0, t ≥ 0.
4
10. Show that a non-empty subset A of n is bounded ⇔ there exists a
real number K such that for each x = (x1 , x2 , . . . , xn ) in A we have
|xi | ≤ K for each subscript i.
11. Show that the real linear space C([−1, 1]) equipped with the norm
given by
1
||x||1 = |x(t)|dt,
−1
ρ(x, y) = ρ(x − y, 0)
and ρ(αx, 0) = |α|ρ(x, 0) ∀ x, y ∈ E and α ∈ 4 (+ )
Define ||x|| = ρ(x, 0), x ∈ E. Prove that || · || is a norm on E and
that ρ is the metric induced by the norm || · || on E.
13. Let E be a linear space of all real valued functions defined on
[0, 1] possessing continuous first-order derivatives. Show that ||f || =
|f (0)| + ||f ||∞ is a norm on E that is equivalent to the norm
||f ||∞ + ||f ||∞ .
2.1.6 Lemma
Interchanging x with y,
||y|| − ||x|| ≤ ||y − x||.
Hence | ||x|| − ||y|| | ≤ ||x − y||.
2.1.7 Lemma
2.1.8 Corollary
4+
Let E be a complete normed linear space over ( ). If {xn }, {yn } ⊂ E,
4+
αn ∈ ( ) and xn → x, yn → y respectively as n → ∞ and αn ∈ ( ) 4+
then (i) xn + yn → x + y (ii) αn xn → x as x → α.
Hence xn + yn → x + y as n → ∞.
||αn xn − αx|| ≤ ||αn (xn − x)+(αn − α)x|| ≤ |α| ||xn − x|| + |αn − α| ||x|| → 0
because {αn } being a convergent sequence is bounded and ||x|| is finite.
∞
In this case we write s = xn . {xn } is said to be absolutely summable
n=1
∞
if ||xn || < ∞.
n=1
Normed Linear Spaces 69
2.1.10 Theorem
K
i.e., for each
> 0 ∃ a K such that ||xn || <
, i.e.,
n=1
0 n 0
0 0
n
0 0
||sn − sm || = 0 xk 0 ≤ ||xk ||
0 0
k=m+1 k=m+1
∞
≤ ||xk ||, ≤
, n, m > K.
n=K
n
In the above, sn = ||xk ||.
k=1
k
1
(a) yn = xnk+1 , (b) ||yk || < , k ≥ 1.
n=0
2k
∞
∞
1
Thus ||yk || ≤ ||y0 || + = ||y0 || + 1 < ∞.
2k
k=0 k=1
Since normed linear spaces can be treated as metric spaces, all concepts
introduced in metric spaces (e.g., balls, spheres, bounded set, separability,
compactness, linear dependence of elements, linear subspace, etc.) have
similar meanings in normed linear spaces. Therefore, theorems proved
in metric spaces using such concepts can have parallels in normed linear
spaces.
(i) The set {x ∈ E : ||x − x0 || < r}, denoted by B(x0 , r), is called the
open ball with centre x0 and radius r.
(ii) The set {x ∈ E : ||x − x0 || ≤ r}, denoted by B(x0 , r), is called a
closed ball with centre x0 and radius r.
(iii) The set {x ∈ E : ||x − x0 || = r}, denoted by S(x0 , r), is called a
sphere with centre x0 and radius r.
Note 2.1.1.
3. Given r > 0
1 0x0 2
0 0
B(0, r) = {x ∈ E : ||x|| < r} = x ∈ E :0 0<1
r
x
= {ry ∈ E : ||y|| < 1} where y =
r
= rB(0, 1).
y = x1 + (1 − λ)x2 , x1 , x2 ∈ X, 0 ≤ λ ≤ 1
x1 , x2 ∈ X ⇒ λx1 + (1 − λ)x2 , x1 , x2 ∈ E, 0 ≤ λ ≤ 1.
Note 2.1.2
(i) For any point x = θ, a ball of radius r > ||x|| with its centre in the
origin, contains the point x.
(ii) Any ball of radius r < ||x|| with centre in the origin does not contain
this point.
(n)
In order to have geometrical interpretations of different abstract spaces lp ,
the n-dimensional pth summable spaces, we draw the shapes of unit balls
for different values of p.
Problems
1. Show that for the norms in examples (ii), (iii), (iv) and (v) the unit
spheres reflect what is shown in figure below:
||x||∞ = 1
||x||4 = 1
||x||2 = 1
||x||1 = 1
Fig. 2.2
ξ2
1
ξ1
−1 1
−1
Fig. 2.3
Since the normed linear space E is a special case of linear space, all
the concepts introduced in a linear space (e.g., linear dependence and
independence of elements, linear subspace, decomposition of E into direct
sums, etc.) have a relevance for E.
Definition: subspace
A set X of a normed linear space E is called a subspace if
Examples
1
||xn − x|| <
, ∀ n,
n
⇒ lim xn = x in X
n→∞
⇒ {xn } is a Cauchy sequence in E and therefore in X.
On the other hand, let X be closed, in which case it contains all of its
limiting points. Hence every Cauchy sequence will converge to some point
in X. Otherwise the subspace X will not be closed.
Examples:
x = (ξ1 , ξ2 , . . . , ξn , 0, . . .) in 4(+),
where ξn = 0 for only finite values of n. Clearly, Φ̂ ⊂ c0 ⊂ l∞ and Φ̂ = c0 .
It may be noted that c0 is the closure of Φ̂ in (l∞ , || · ||∞ ). Thus Φ̂ is not
closed in l∞ and hence Φ̂ is an incomplete normed linear space equipped
with the norm induced by the norm || · ||∞ on l∞ .
Φ̂ ⊂ lp ⊂ c0
Problems
+ +
2. If n ≥ m ≥ 0, prove that n ({a, b]) ⊂ m ([a, b]) and that the space
+ n
+
([a, b]) with the norm induced by the norm on m ([a, b]) is not
closed.
Although infinite dimensional normed linear spaces are more general than
finite dimensional normed linear spaces, finite dimensional normed linear
spaces are more useful. This is because in application areas we consider
the finite dimensional spaces as subspaces of infinite dimensional spaces.
Quite a number of interesting results can be derived in the case of finite
dimensional spaces.
2.3.1 Theorem
y = η1 e1 + η2 e2 + · · · + ηn en .
Now, since the unit ball S(0, 1) in En is compact and the function
f (ξ1 , ξ2 , . . . , ξn ) defined on it is continuous, it follows that f (ξ1 , . . . , ξn ) has
a minimum on S. Hence,
or in other words
||x − y|| ≥ γ||x̃ − ỹ||. (2.8)
From (2.7) and (2.8) it follows that the mapping of E onto En is one-
to-one.
The mapping from E onto En is one-to-one and onto. Both the mapping
and its inverse are continuous. Thus, the mapping is a homeomorphism.
The homeomorphism between E and En implies that in a finite dimensional
78 A First Course in Functional Analysis
||β1 e1 + β2 e2 + · · · + βn en || ≥ c (2.10)
n
Note that |βi | = 1.
i=1
Hence it suffices to prove the existence of a c > 0 such that (2.10) holds
n
for every n-tuples of scalars β1 , β2 , . . . , βn with |βi | = 1.
i=1
Suppose that this is false. Then there exists a sequence {ym } of vector
n
(m)
(m)
(m)
ym = β1 e1 + · · · + βn en , |βi | = 1
i=1
n
(m) (m)
Since |βi | = 1 , we have |βi | ≤ 1. Hence for each fixed i the
i=1
sequence
(m) (1) (2)
(βi ) = (βi , βi , . . .)
(m)
is bounded. Consequently, by the Bolzano-Weierstrass theorem, {βi } has
a convergent subsequence. Let βi denote the limit of the subsequence and
let {y1,m } denote the corresponding subsequence of {ym }. By the same
Normed Linear Spaces 79
n
(m)
yn,m = γi ei ,
i=1
n
(m) (m)
|γi | = 1 with scalars γi → βi as m → ∞
i=1
n
yn,m → y = βi ei
i=1
where |βi | = 1, so that not all βi can be zero.
i
(m)
4+
This shows that {αi } is a Cauchy sequence in ( ) for i = 1, 2, . . . , n.
Hence the sequence converges. Let αi denote the limit. Using these n limits
α1 , α2 , . . . , αn , let us construct y as
y = α1 e1 + α2 e2 + · · · + αn en
Here y ∈ X and
0 n 0
0 0 n
0 (m) 0 (m)
||ym − y|| = 0 (αi − αi )ei 0 ≤ |αi − αi | ||ei ||.
0 0
i=1 i=1
(m)
Since αi → αi as m → ∞ for each i, ym → y as m → ∞. This shows
that {ym } is convergent in X. Since {ym } is an arbitrary Cauchy sequence
in X it follows that X is complete.
Definition 2: Two norms ||·|| and ||·|| on the same linear space E are said to
be equivalent norms on E if the identity mapping IE : (E, || · ||) → (E, || · || )
is a topological homoeomorphism of (E, || · ||) onto (E, || · || ).
Theorem: Two norms || · || and || · || on the same normed linear space
Normed Linear Spaces 81
x = α1 e1 + α2 e2 + · · · + αn en . (2.11)
Then by lemma 2.3.2, we can find a constant c > 0 such that
0 0 n
0n 0
0 0
||x|| = 0 αi ei 0 ≥ c |αi | (2.12)
0 0
i=1 i=1
On the other hand,
0 n 0
0 0
0 0
||x|| = 0 αi e i 0
0 0
i=1
n
≤ |αi | ||ei || (2.13)
i=1
82 A First Course in Functional Analysis
n
≤ k1 |αi |
i=1
where k1 = max ||ei || .
Hence ||x|| < α2 ||x|| (2.14)
k1
where α2 =
c
Interchanging ||x|| and ||x|| we obtain as in the above
||x|| ≥ α1 ||x||
−1
k2
where α1 = , k2 = max ||ei ||.
c
Problems
Then show that || ||, ||p||1 , ||p|||∞ are norms on E, ||p|| ≤ ||p||1 and
||p||∞ ≤ ||p||1 for all p ∈ E.
12. Show that equivalent norms on a linear space E induce the same
topology for E.
13. If two norms || · || and || · ||0 on a linear space are equivalent, show
that (i) ||xn − x|| → 0 implies (ii) ||xn − x||0 → 0 (and vice versa).
Proof: Take any v ∈ Z − Y and denote its distance from Y by d (fig. 2.4),
v
d y0
Fig. 2.4
84 A First Course in Functional Analysis
Clearly, d > 0 since Y is closed. We now take any θ ∈ (0, 1). By the
definition of an infinum there is a y0 ∈ Y such that
d
d ≤ ||v − y0 || ≤ (2.15)
θ
d
(note that θ > d since 0 < θ < 1).
1
Let, z = c(v − y0 ) where c = .
||v − y0 ||
Then ||z|| = 1 and we shall show that ||z − y|| ≥ θ for every y ∈ Y .
d d
||z − y|| = c||v − y1 || ≥ cd = ≥ d = θ.
||v − y0 || θ
2.3.8 Lemma
1
||y n || = 1 and dist(y n , Zn ) ≥ .
2
Problems
2.4.1 Theorem
Let E be a normed linear space over 4(+) and let L be a closed subspace
of E.
Proof: We first show that || · ||q defines a norm on E/L. We note first that
||x + L||q ≥ 0, ∀ x ∈ E.
Now x ∈ L as L is closed.
Hence x + L = L
1
xnk + L = yk + L and ||yk+1 − yk || < (k = 1, 2, . . .)
2k
Then for p = 1, 2, . . .
p
p
1 1
||yk+p − yk || ||yk+i − yk+i−1 || < = k+p−1
i=1 i=1
2k+i−1 2
Hence the Cauchy sequence {xn + L} converges in E/L and thus E/L is
complete.
Problems
˜ then
˜ in Ẽ,
{xn } → x̃ in Ẽ and xn → x̃
Example: The completion of the normed linear space (P[a, b], || · ||∞ ) where
P[a, b] is the set of all polynomials with real coefficients defined on the
closed interval [a, b], is the space (C([a, b]), || · ||∞ ).
CHAPTER 3
HILBERT SPACE
91
92 A First Course in Functional Analysis
Hence inner product spaces are normed linear spaces, and Hilbert
spaces are Banach spaces.
In (c) the bar denotes complex conjugation. In case,
The proof that (3.1) satisfies the axioms (a) to (d) of a norm [see 2.1] will
be given in section 3.2.
From (a) to (d) we obtain the formula,
⎫
(a ) αx + βy, z = αx, z + βy, z for all scalars α, β. ⎪
⎬
(b ) x, αy = αx, y (3.4)
⎪
⎭
(c ) x, αy + βz = αx, y + βx, z.
3.1.2 Observation
It follows from (a ) that the inner product is linear in the first argument,
while (b ) shows that the inner product is conjugate linear in the second
argument. Consequently, the inner product is sesquilinear, which means
that 1 12 times linear.
3.1.3 Examples
1. 4 n
, n dimensional Euclidean space
4
The space n is a Hilbert space with inner product defined by
n
x, y = ξi ηi (3.5)
i=1
Hilbert Space 93
∞
∞
1/2 ∞
1/2
2 2
We have x, y = ξi ηi ≤ |ξi | |ηi | <∞ (3.8)
i=1 i=1 i=1
From (3.7) we obtain the norm defined by
∞
12
1
x = x, x 2 = |ξi |2 .
i=1
Using the metric induced by the norm, for l2 , we see that all the axioms
3.1.1(a)–(d) are fulfilled.
4. Space L2 ([a, b])
The inner product is defined
b
x, y = x(t)y(t)dt (3.9)
a
for x(t), y(t) ∈ L2 ([a, b]) i.e., x(t), y(t) are Riemann square integrable.
The norm is then
12
b
1
2
x = x, x 2 = x(t) dt where x(t) ∈ L2 ([a, b]).
a
Using the metric induced by the norm we can show that L2 ([a, b]) is a
Hilbert space.
In the above x(t) is a real-valued function.
In case x(t) and y(t) are complex-valued functions we can define the
b
inner product x, y as x, y = x(t)y(t)dt with the norm given by
a
12
b
2
x(t) = |x(t)| dt because x(t)x(t) = |x(t)|2 .
a
x
Fig. 3.1 Parallelogram with sides x and y in the plane
Now, in ,4
1 1
(x + y2 − x − y2 ) = [x + y, x + y − x − y, x − y]
4 4
1
= [x,x + y,y + x, y + y, x − x,
x − y,
y + x, y + y, x].
4
4
= x, y since in , x, y = y, x.
In +
1
[(x + y2 − x − y2 ) + i(x + iy2 − x − iy2 )]
4
1
= [{x, x + y,
y + x, y + y, x − x,
x − y,
y + x, y + y, x}
4
y − x,
x + x, iy + iy, x + iiy,
+ i{x, y + x, iy + iy, x}]
x − iiy,
1
= [2x, y + 2x, y + 2x, y − 2x, y] = x, y.
4
which is the polarization identity.
3.3.4 Theorem
A norm · on a linear space E is induced by an inner product , on it
if and only the norm satisfies the parallelogram law. In that case the inner
product , is given by the polarization equality.
Proof: Suppose that the norm is induced by the inner product. Then
the parallelogram law holds true. Furthermore the inner product can be
recovered from the norm by the polarization equality.
Conversely, let us suppose that · obeys the parallelogram law and
, is defined by the polarization equality as given in (3.11). We have to
show that , is an inner product and generalize · on E. Let us consider
the formula (3.11) for the complex space.
1
= [(4x2 + 2i(x|2 − x2 )]
4
= x2
Therefore inner product , generates the norm · .
(ii) For all x, y ∈ E we have,
1
y, x = [y + x2 − y − x2 − i(y + ix2 − y − ix2 )]
4
1
= [x + y2 − x − y2 + i(x + iy2 − x − iy2 )] = x, y.
4
(iii) Let u, v, w ∈ X. Then parallelogram law yields
3
(u + v) + w2 + (u + v) − w2 = 2(u + v2 + w2 )
(u − v) + w2 + (u − v) − w2 = 2(u − v2 + w2 ).
On substraction we get,
(iv) Next we want to prove condition (b), i.e., αx, y = αx, y, for
every complex scalar α and ∀ x, y ∈ E. We shall prove it in stages.
Stage 1. Let λ = m, a positive integer, m > 1.
9x : 1
Hence , y = x, y.
n n
If m is a negative integer, splitting m as (m − 1) + 1 we can show that (b)
is true.
m
Stage 2. Let α = r = be a rational number, m and n be prime to each
n
other.
m 9x : 9m :
Then rx, y = x, y = m ,y = x, y = rx, y.
n n n
Stage 3. Let α be a real number. Then there exists a sequence {rn } of
rational numbers, such that rn → α as n → ∞.
Thus we have shown that , is the inner product inducing the norm ·
on E.
3.3.5 Lemma
Let E be an inner product space with an inner product , .
(i) The linear space E is uniformly convex in the norm · , that is,
for every
> 0, there is some δ > 0, such that for all x, y ∈ E with
x ≤ 1, y ≤ 1 and x − y ≤
, we have x + y ≤ 2 − 2δ.
(ii) The scalar product is a continuous function with respect to norm
convergence.
Proof: (i) Let
> 0. Given x, y ∈ E with x ≤ 1, y ≤ 1 and x−y ≥
.
Then
≤ x − y ≤ x + y ≤ 2.
The parallelogram law gives
xn , yn −→ x, y as n −→ ∞.
3.4 Orthogonality
3.4.1 Definitions (orthogonal, acute, obtuse)
Let x, y be vectors in an inner product space.
(i) Orthogonal: x is said to be orthogonal to y or written as x ⊥ y if
x, y = 0.
(ii) Acute: x is said to be acute to y if x, y ≥ 0.
(iii) Obtuse: x is said to be obtuse to y if x, y ≤ 0.
4 4
In 2 or 3 , x is orthogonal to y if the angle between the vectors is 90◦ .
Similarly when x is acute to y, the angle between x and y is less than or
equal to 90◦ . We can similarly explain when x, y ≤ 0 the angle between x
and y is greater than or equal to 90◦ . This geometrical interpretation can
be extended to infinite dimensions in an inner product space.
3.4.2 Definition: subspace
A non-empty subset X of the inner product space E is said to be a
subspace of E if
(i) X is a (linear) subspace of E considered as a linear space.
(ii) X admits of a inner product , X induced by the inner product
, on E, i.e.,
x, yX = x, y ∀ x, y ∈ E.
Note 3.4.1. A subspace X of an inner product E is itself an inner product
space and the induced norm · X on X coincides with the induced norm
· on E.
Hilbert Space 101
Re (x, y)
θx,y = arc cos + , 0 ≤ θx,y ≤ π
(x, x)(y, y)
4
5. [Limaye [33]] Let X be a normed space over . Show that the norm
satisfies the parallelogram law, if and only if in every plane through
the origin, the set of all elements having norm equal to 1 forms an
ellipse with its centre at the origin.
6. Let {xn } be a sequence in a Hilbert space H and x ∈ H such that
lim xn = x, and lim xn , x = x, x. Show that lim xn = x.
n→∞ n→∞ n→∞
7. Let C be a convex set in a Hilbert space H, and d = inf{x, x ∈ C}.
If {xn } is a sequence in C such that lim xn = d, show that {xn }
n→∞
is a Cauchy sequence.
8. (Pythagorean theorem) (Kreyszig [30]). If x ⊥ y is an inner
product on E, show that (fig. 3.2),
y
y x+ y
x
Fig. 3.2
x x A
∧ ∧ ∧
No y (A unique y ) (infinitely many y ’s)
Fig. 3.3 Fig. 3.4 Fig. 3.5
x = y + z, (3.17)
|x − yn , h|2
or ≤ (dn − d).
h2
√
or |x − yn , h| ≤ h dn − d. (3.18)
Inequality (3.18) is evidently satisfied for h = 0.
It then follows that
|ym − yn , h| ≤ |x − yn , h| + |x − ym , h|
√ √
≤ ( dn − d + dm − d)h (3.19)
(3.19) yields
|yn − ym , h| + +
≤ ( dn − d + dm − d)
h
Taking supremum of LHS we obtain,
√ √
yn − ym ≤ ( dn − d + dm − d) (3.20)
x − yn , h = 0 (3.21)
x = y + z (3.23)
y − y = z − z. (3.24)
O L
B
Fig. 3.6 Projection of x on L
Proof: Let d = inf x0 − y2 . Let us choose {yn } in L such that
y∈L
lim x − yn 2 = d (3.25)
n→∞
ym + y n
or ym − yn 2 = 2(ym − x2 + yn − x2 ) − 4 − x2 .
2
ym + yn
Since L is convex, ym , yn ∈ L ⇒ ∈L
2
ym + yn
Hence − x2 ≥ d.
2
Hence, ym − yn 2 ≤ 2ym − x2 + 2yn − x2 − 4d.
Using (3.25) we conclude from above that {yn } is Cauchy.
Since L is closed, {yn } → y0 (say) ∈ L as n → ∞.
Then x − y0 2 = d.
For uniqueness, let us suppose that x − z0 2 = d, where z0 ∈ L.
Then, y0 − z0 2 = (y0 − x) − (z0 − x)2
= 2y0 − x2 + 2z0 − x2 − (y0 − x) + (z0 − x)2
0 02
0 y0 + z0 0
= 4d − 4 0
0 2 − x 0 ≤ 4d − 4d = 0
0
Hence y0 = z0 .
3.5.4 Theorem
Let H be a Hilbert space and L be a closed convex set in H and
x ∈ H −L. Then there is a unique y0 ∈ L (i) such that x−y0 = inf x−y
y∈L
and (ii) x − y0 ∈ L⊥ .
Theorem 3.5.3 guarantees the existence of a unique y0 ∈ L such that
x − y0 = inf x − y.
y∈L
If y ∈ L and
is a scalar, then y0 +
y ∈ L, so that
3.5.5 Lemma
In order that a linear subspace L is everywhere dense in a Hilbert space
H, it is necessary and sufficient that an element exists which is different
from zero and orthogonal to all elements of M .
Proof: Since M is everywhere dense in H, x ⊥ M implies that x ⊥ M . By
hypothesis M = H and consequently, x ⊥ H, in particular x ⊥ x implying
x = θ.
Conversely, let us suppose M is not everywhere dense in H. Then
M = H and there is an x ∈ M and x ∈ H. By theorem 3.5.1 x = y + z,
y ∈ M , z ⊥ M , and since x ∈ M , it follows that z = θ, which is a
contradiction to our hypothesis. Hence M = H.
x = x1 + x2 + · · · + xn , xi ∈ Xi .
Hence {zn } is Cauchy and {xn } and {yn } are Cauchy sequences in L
and L⊥ respectively. Since L and L⊥ are closed subspaces of H, they are
complete. Hence {xn } → x an element in L and {yn } → y, an element in
L⊥ as n → ∞.
x − y0 = inf x − y = d > 0
y∈L+L⊥
Hilbert Space 109
Since z = x − y = x − P x = (I − P )x.
Thus PH = L (I − P )H = L⊥
Now, P y = y, y ∈ L and P z = 0, z ∈ L⊥ .
Thus the range of P and its null space are mutually orthogonal. Hence the
projection mapping P is called an orthogonal projection.
3.6.3 Theorem
A subspace L of a Hilbert space H is closed if and only if L = L⊥⊥ .
If L = L⊥⊥ , then L is a closed subspace of H, because L⊥⊥ is already
a closed subspace of H [see note 3.6.1].
Conversely let us suppose that L is a closed subspace of H. For any
subset L of H, we have L ⊆ L⊥⊥ [see note 3.6.1]. So it remains to prove
that L⊥⊥ ⊆ L.
Let x ∈ L⊥⊥ . By projection theorem, x = y + z, y ∈ L and z ∈ L⊥ .
Since L ⊆ L⊥⊥ , it follows that y ∈ L⊥⊥ .
L⊥⊥ being a subspace of H, z = x − y ∈ L⊥⊥ .
Hence z ∈ L⊥ ∩ L⊥⊥ . As such z ⊥ z, i.e., z = θ. Hence, z = x − y = θ.
Thus x ∈ L. Hence L⊥⊥ ⊆ L. This proves the theorem.
110 A First Course in Functional Analysis
3.6.4 Theorem
Let L be a non-empty subset of a Hilbert space H. Then, span L is
dense in H if and only if L⊥ = {θ}.
Proof: We assume that span L is dense in H. Let M = span L so that
M = H. Now, {θ} ⊂ L⊥ . Let x ∈ L⊥ and since x ∈ H = M , there exists
a sequence {xn } ⊆ M such that
lim xn = x.
n→∞
O x
e1
Fig. 3.7 Expressing a vector in terms of two perpendicular vectors
a1 + a2 + · · · + ak 2 = a1 + a2 + · · · + ak , a1 + a2 + · · · + ak
k
k
k
k
= ai , aj = ai , ai = ai 2 .
i=1 j=1 i=1 i=1
2π
vn , vm = cos nt cos mt dt
⎧0
⎨ 0, m = n
= π, m = n = 1, 2 · · ·
⎩
2π, m = n = 0.
√
Hence vn = π for n = 0.
1 cos t cos nt
Therefore, √ , √ , √ is an orthonormal sequence.
2π π π
3.7.7 Examples
1. (Orthonormal polynomials) Let L2,ρ ([a, b]) be the space
of square-summable functions with weight functions ρ(t). Let us
take a linearly independent set 1, t, t2 · · · tn · · · in L2,ρ ([a, b]). If we
orthonormalize the above linearly independent set, we get Chebyshev
system of polynomials, p0 = const., p1 (t), p2 (t), . . . , pn (t), . . . which are
orthonormal with weight ρ(t), i.e.,
b
ρ(t)pi (t)pj (t)dt = δij .
a
1 1
= (v n )(n−1) (v n )(n) − (vn )(n−1) (v n )(n+1) dt.
−1 −1
1 1
= (−1)n (2n)! vn dt = 2(2n)! (1 − t2 )n dt
−1 −1
π
2
= 2(2n)! cos2n+1 α dα (t = sin α)
0
22n+1 (n!)2
= .
2n + 1
0; 0
0 2n + 1 0
0 0
Thus 0 Pn 0 = 1.
0 2 0
1
1
Next, Pm , Pn = m+n ((t2 − 1)m )(m) ((t2 − 1)n )(n) dt
2 m!n! −1
1
1
= m+n (v m )(m) (v n )(n) dt
2 m!n! −1
where m > n (suppose) and v = t2 − 1.
+1
1
= m+n (v m )(m) (v n )(n−1)
2 m!n! −1
1
−2m (v m )(m−1) (vn )(n+1) dt
−1
1
(−1)n 2m · · · 2(m − n)
= (v m )(m−n) dt = 0 m > n.
2m+n m!n! −1
A similar conclusion is drawn if n > m.
and (3.32) can also be obtained by applying the power series method to
(3.34).
Hilbert Space 117
y
P0
1
P1
P2
x
−1 1
−1
<
1 2
− t2
We want to show that en (t) = √ e Hn (t), n ≥ 1 (3.35)
2n n! π
t2
e− 2
and e0 (t) = √ 1 H0 (t)
( π) 2
dn −t2 2
where H0 (t) = 1, Hn (t) = (−1)n et (e ), n = 1, 2, 3, . . . (3.36)
dtn
Hn are called Hermite polynomials of order n. H0 (t) = 1.
Performing the differentiations indicated in (3.36) we obtain
N
2n−2j
Hn (t) = n! (−1)j tn−2j (3.37)
j=0
j!(n − 2j)!
n (n − 1)
where N = if n is even and N = if n is odd. The above form
2 2
can also be written as
N
(−1)j
Hn (t) = n(n − 1) · · · (n − 2j + 1)(2t)n−2j (3.38)
j=0
j!
t2
e− 2
Thus e0 (t) = √ 1 H0 (t)
( π) 2
√ − t2 t2
<
2te 2 e− 2 (2t) 1 t2
e1 (t) = √ 1 = √ 1 = n
√ e− 2 Hn (t)n=1 .
( π) 2 (2 π) 2 2 n! π
(3.36) yields explicit expressions for Hn (t) as given below for a few values
of n.
H0 (t) = 1 H1 (t) = 2t
H2 (t) = 4t2 − 2 H3 (t) = 8t3 − 12t
H4 (t) = 16t4 − 48t2 + 12 H5 (t) = 32t5 − 160t3 + 120t.
We next want to show that {en (t)} is orthonormal where en (t) is given
by (3.35) and Hn (t) by (3.37),
< ∞
1 2
en (t), em (t) = n+m
√ 2
e−t Hm Hn dt.
2 n!m!( π) −∞
Differentiating (3.38) we obtain for n ≥ 1,
M
(−1)j
Hn (t) = 2n (n − 1)(n − 2) · · · (n − 2j)(2t)n−1−2j
j=0
j!
= 2nHn−1 (t)
(n − 2) (n − 1)
where N = if n is even and N = if N is odd.
2 2
Hilbert Space 119
2
Let us assume m ≤ n and u = e−t . Integrating m times by parts we
obtain from the following integral,
∞
2
(−1) n
e−t Hm (t)Hn (t)dt
−∞
∞
= Hm (t)(u)(n) dt
−∞
∞ ∞
= Hm (t)(u)(n−1) − 2m Hm−1 (u)n−1 dt
−∞ −∞
∞
= −2m Hm−1 (u)(n−1) dt
−∞
= ······
∞
= (−1)m 2m m! H0 (t)(u)(n−m) dt
−∞
∞
= (−1)m 2m m!(u)(n−m−1)
−∞
=0 if n > m because as t → ±∞, t2 → ∞ and u → 0.
This proves orthogonality of {em (t)}, when m = n
∞ ∞
2 2
(−1)n e−t Hn2 (t)dt = (−1)n 2n n! H0 (t)e−t dt
−∞ −∞
n n √
= (−1) 2 n! π.
∞
2 √
Hence e−t Hn2 (t)dt = 2n n! π.
−∞
This proves orthonormality of {en }.
− 2t − 2t
∞
−t te−t ∞ ∞
te ,e = te dt = + e−t dt = 1.
0 −1 0 0
t
g1 = (t − 1)e− 2
∞
t t
g1 2 = (t − 1)e− 2 , (t − 1)e− 2 = (t − 1)2 e−t dt = 1.
0
− 2t
e1 (t) = (t − 1)e .
− 2t
Let us take en (t) = e Ln (t), n = 0, 1, 2 (3.39)
where the Laguerre polynomials of order n is defined by
et dn n −t
L0 (t) = 1, Ln (t) = (t e ), n = 1, 2 (3.40)
n! dtn
(−1)j
n
n
i.e. Ln (t) = − tj (3.41)
j! j
j=0
L0 (t) = 1, L1 (t) = 1 − t
1 3 1
L2 (t) = 1 − 2t + t2 L3 (t) = 1 − 3t + t2 − t3 .
2 2 6
2 2 3 1 4
L4 (t) = 1 − 4t + 3t − t + t
3 24
The Laguerre polynomials Ln are solutions of the Laguerre differential
equations
tLn + (1 − t)Ln + nLn = 0. (3.42)
In what follows we find out e2 (t) by Gram-Schmidt process.
∞
= (t4 − 8t3 + 20t2 − 16t + 4)e−t dt
0
= 4.
! "
e2 (t) = 1 − 2t + 12 t2 e−t .
The orthogonality {Lm (t)} can be proved as before.
n
n
n
n
= x|2 − αi x, ei − αi ei , x + αi αj ei , ej
i=1 i=1 i=1 j=1
n
n
n
= x2 − αi di − α i di + |αi |2 where di = x, ei .
i=1 i=1 i=1
n
n
Therefore, x − αi ei 2 = x2 − |di |2 + |αi − di |2 .
i=1 i=1
∞
∞
Then y= di ei where di = y, ei = x, ei and |di |2 = y2 .
i=1 i=1
set. Therefore, C =
Φ. Now let us consider any totally ordered subfamily
in C. The union of sets in this subfamily is clearly an orthonormal set
and is an upper bound for the totally ordered subfamily. Therefore, by
Zorn’s lemma (1.1.4), we conclude that C has a maximal element which is
a complete orthonormal set in H.
The next theorem provides another characterization of a complete
orthonormal system.
3.8.4 Theorem
Let {ei } be an orthonormal set in a Hilbert space H. Then
{ei } is a complete orthonormal set if and only if it is impossible
to adjoin an additional element e ∈ H, e = θ to {ei } such that
{ei , e} is an orthonormal set in H.
Proof: Suppose {ei } is a complete orthonormal set. Let it be possible to
adjoin an additional vector e ∈ H of unit norm e = θ, such that {e, ei } is
an orthonormal set in H i.e. e ⊥ {ei }, e is a non-zero vector of unit norm.
But this contradicts the fact that {ei } is a complete orthonormal set. On
the other hand, let us suppose that it is impossible to adjoin an additional
element e ∈ H, e = θ, to {ei } such that {e, ei } is an orthonormal set.
Or, in other words, there exists no non-zero e ∈ H of unit norm such that
e ⊥ {ei }. Hence the system {ei } is a complete orthonormal system.
In what follows we define a closed orthonormal system in a Hilbert space
and show that it is the same as a complete orthonormal system.
3.8.5 Definition: closed orthonormal system
An orthonormal system {ei } in H is said to be closed if the subspace
L spanned by the system coincides with H.
3.8.6 Theorem
A Fourier series with respect to a closed orthonormal system,
constructed for any x ∈ H, converges to this element and for
124 A First Course in Functional Analysis
holds.
Proof: Let {ei } be a closed orthonormal system. Then the subspace
spanned by {ei } coincides with H.
Let x ∈ H be any element. Then the Fourier series (c.f. 3.7.8) WRT
the closed system is given by
∞
x= di e i where di = x, ei ,
i=1
Then by the relation (3.43), we have,
∞
x2 = d2i where di = x, ei and ei 2 = 1.
i=1
3.8.7 Corollary
An orthonormal system is complete if and only if the system
is closed.
Let {ei } be a complete orthonomal system in H. If {ei } is not closed,
let the subspace spanned by {ei } be L where L = H. If any non-zero x is
not orthogonal to L, then x is also not orthogonal to L. Thus Φ ≡ H − L.
This contradicts that {ei } is not closed.
Conversely, let us suppose that the orthonormal system {ei } is closed.
Then, by theorem 3.8.6, we have for any x ∈ H,
∞
x2 = d2i where di = x, ei .
i=1
One can conclude that the partial sums of 3.48 is the vector un (t) =
n
√1 xk eikt converges to the vector x in the sense of L2 , i.e.,
2π
k=−∞
1 2
The above relation shows that eβ is a linear combination of eλ |λ∈Λ ,
λ=β
which is a contradiction, since {eλ |λ∈Λ } is an orthonormal set and hence
linearly independent.
Hence {eλ |λ∈Λ } cannot be non-enumerable.
Problems [3.7 and 3.8]
1. Let {a1 , a2 , . . . an } be an orthogonal set in a Hilbert space H,
and α1 , α2 , . . . αn be scalars such that their absolute values are
respectively 1. Show that
α1 a1 + · · · + αn an = a1 + a2 + · · · + an .
2. Let {en } be an orthonormal sequence in a Hilbert space H. If {αn }
∞
be a sequence of scalars such that |αi |2 converges then show that
i=1
∞
αi ei converges to an x
H and αn = x, en , ∀ n ∈ N.
i=1
126 A First Course in Functional Analysis
5. Show that on the unit disk B(|z| < 1) of the complex plane z = x+iy,
the functions
12
k
gk (x) = z k−1 (k = 1, 2, 3)
π
form an orthonormal system under the usual definition of a scalar
product in a complex plane.
6. Let {eα |λ ∈ Λ} be an orthonormal set in a Hilbert space H.
(a)
φ(0) (x) = 1 x ∈ [0, 1]
⎧ ? 1
⎪
⎪ 1, x ∈ 0,
⎪
⎪
⎪
⎨ 2
1 )
(0)
φ1 (x) = −1, x ∈ ,1
⎪
⎪ 2
⎪
⎪
⎪
⎩ 0, x = 1
2
Hilbert Space 127
and for m = 1, 2, . . . ; K = 1, . . . 2m
⎡
√ K − 1 K − 12
⎢ 2 , x∈
m
m
, m
⎢
⎢ 2 1 2
√ K−2 K
φm (x) = ⎢
(K)
⎢ − 2m , x ∈ ,
⎢ 2m 2m
⎣ K −1 K
0, x ∈ [0, 1] ∼ ,
2m 2m
(K)
and at that finite set of points at which φm (x) has not yet been
(K)
defined, let φm (x) be the average of the left and right limits of
(K) (K)
φm (x) as x approaches the point in question. At 0 and 1, let φm (x)
assume the value of the one-sided limit. Show that the Haar system
given by
(0) (0) (K)
{φ0 , φ1 · · · φm , m = 1, 2, . . . ; K = 1, . . . 2m }
is orthonormal in L2 ([0, 1]).
9. If H has a denumerable orthonormal basis, show that every
orthonormal basis for H is denumerable.
10. Show that the Legendre differential equation can be written as
[(1 − t2 )Pn ] = −n(n + 1)Pn .
Multiply the above equation by Pn and the corresponding equation
in Pm by Pn . Then subtracting the two and integrating resulting
equation from −1 to 1, show that {Pn } in an orthogonal sequence in
L2 ([−1, 1]).
11. (Generating function) Show that
∞
1
√ = Pn (t)w n .
1 − 2tw + w2 n=0
where the subscripts H and l2 denote the respective spaces whose norms
are taken.
Moreover, it is clear that if x ∈ H corresponds to x̃ ∈ l2 , and y ∈ H
corresponds to ỹ ∈ l2 , then x ± y corresponds to x̃ ± ỹ. It then follows from
(3.49) that
x − yM = x̃ − ỹl2 . (3.50)
Let us suppose that z = {ζi } is an arbitrary element in l2 . We next consider
n
in H the elements zn = ζi ei , n = 1, 2, . . .
i=1
n
n
We have then zn − zm 2 = ζi ei 2 = |ζi |2 .
i=m+1 i=m+1
Now, zm − zn → 0 as n, m → ∞.
scalar λ. Since x ± yH = x̃ ± ỹl2 and λxH = λx̃l2 , it follows that
the correspondence between H and l2 is both isometric and isomorphic.
We thus obtain the following theorem.
3.9.1 Theorem
Every complex (real) separable Hilbert space is isomorphic and isometric
to a complex (real) space l2 . Hence all complex (real) separable Hilbert
spaces are isomorphic and isometric to each other.
CHAPTER 4
LINEAR OPERATORS
130
Linear Operators 131
x = (ξ1 , ξ2 , . . . , ξn ) ∈ Ex
If x1 = (ξ11 , ξ21 , . . . , ξn1 ) and x2 = (ξ12 , ξ22 , . . . , ξn2 ) and y 1 =
(η11 , η21 , . . . , ηn1 )
and y 2 = (η12 , η22 , . . . , ηn2 ) such that
n
Ax1 = y1 ⇒ aij ξj1 = ηi1 , i = 1, 2, . . . , n
j=1
n
2 2
Ax = y ⇒ aij ξj2 = ηi2 , i = 1, 2, . . . , n
j=1
Then A(x1 + x2 ) = ( aij (ξj1 + ξj2 )) = ( aij ξj1 ) + ( aij ξj2 )
j j j
= Ax1 + Ax2 = y 1 + y 2 .
Since the sup of (4.2) exists, sup |y(t)| also exists. Moreover the operator
dx
A= is linear for
dt
x1 , x2 ∈ C 1 ([a, b]) A(x1 + x2 ) = Ax1 + Ax2
⇒
λ∈ ( ) 4+ A(λx) = λAx
4.1.4 Continuity
We know that the continuity of A in the case of a metric space means
that there is a δ > 0 such that the collection of images of elements in the
ball B(x, δ) lie in B(Ax,
).
4.1.1b Example
(m)
Let us suppose in example 4.1.1a {ξi } is convergent.
(m) (p)
n
(m) (p)
Then ηi − ηi = aij (ξj − ξj )
j=1
Hence by Cauchy-Bunyakovsky-Schwartz inequality (1.4.3)
n
(m) n
n
n (m)
(η − η
(p) 2
) ≤ a 2 (p)
(ξi − ξi )2
i i ij
i=1 i=1 j=1 j=1
Linear Operators 133
(m) (m)
where xm = {ξi }, ym = Axm = {ηi }.
n
n
(m)
Since a2ij is finite, convergence of {ξi } implies convergence of
i=1 j=1
(m)
{ηi }. Hence continuity of A is established.
4.1.2b Example
Consider Example 4.1.2a.
Let {xn (t)} converge to x(t) in the sense of convergence in C([0, 1]), i.e.,
converges uniformly in C[(0, 1]). Now in the case of uniform convergence
we can take the limit under the integral sign.
b b
It follows that lim K(t, s)xm (s)ds = K(t, s)x(s)ds i.e.
m a a
lim Axm = Ax and the continuity of A is proved.
m→∞
4.1.3b Example
We refer to the example 4.1.3a. The operator A, in this case, although
additive and homogeneous, is not continuous. This is because the
derivative of a limit element of a uniformly convergent sequence of functions
need not be equal to the limit of the derivative of these functions, even
though all these derivatives exist and are continuous.
4.1.4 Example
Let A be a continuous linear operator. Then
(i) A(θ) = θ (ii) A(−z) = −Az for any z ∈ Ex
Proof: (i) for any x, y, z ∈ Ex , put x = y + z and consequently y = x − z.
Now,
Ax = A(y + z) = Ay + Az = Az + A(x − z)
Hence, A(x − z) = Ax − Az (4.3)
A(−z) = −Az
4.1.5 Theorem
If an additive operator A, mapping a real linear space Ex into a real
linear space Ey s.t. y = Ax, x ∈ Ex , y ∈ Ey , be continuous at a point
x∗ ∈ Ex , then it is continuous on the entire space Ex .
Proof: Let x be any point of Ex and let xn → x as n → ∞. Then
xn − x + x∗ → x∗ as n → ∞.
Since A is continuous at x∗ ,
134 A First Course in Functional Analysis
4.1.6 Theorem
An additive and continuous operator A defined on a real linear space is
homogeneous.
Proof: (i) Let α = m, a positive integer. Then
A(αx) = Ax + Ax + · · · + Ax = mAx
A BC D
m terms
A(si x) = si Ax (4.4)
since A is continuous at αx, α is a real number, and since lim si x = αx
i→∞
we have,
Linear Operators 135
(A + B)x = Ax + Bx
(λA)x = λAx.
(A + B)x = (A − A)x = 0 · x = Ax − Ax = θ
Thus 0, the null operator, is an element of the said space. The limit of
a sequence is defined in a space of linear operators by assuming for example
that An → A if An x → Ax for every x ∈ Ex .
This space of continuous linear operators will be discussed later.
4.1.8 The ring of continuous linear operators
4+
Let E be a linear space over a scalar field ( ). We next consider the
space of continuous linear operators mapping E into itself. Such a space
we denote by (E → E). The product of two linear operators A and B in
(E → E) is denoted by AB = A(B), i.e., (AB)x = A(Bx) for all x ∈ E.
Let xn → x in E. A and B being continuous linear operators,
Axn → Ax and Bxn → Bx.
Since AB is the product of A and B in (E → E),
An+1 = A · An ∈ (E → E)
Thus AB = BA.
pn (A) = a0 I + a1 A + a2 A2 + · · · + an An .
4.1.1c Example
(m) (m)
Now refer back to example 4.1.1a. Let, xm = {ξi }. ym = {ηi }.
Then ym = Axm . Let xm → x as m → ∞. We assume that xm ∈ Ex , the
n-dimensional Euclidean space. Then
n
(m) 2
(ξi ) < ∞.
i=1
Now, by Cauchy-Bunyakovsky-Schwartz’s inequality, (Sec. 1.4.3)
⎧ ⎫
⎨ n n ⎬
(m) (m)
y − ym = {ηi − ηi } = aij (ξj − ξj )
⎩ ⎭
i=1 j=1
⎛ ⎞1/2
n n
= aij (ξj − ξj ) ≤ ⎝
(m) (m)
or |ηi − ηi | a2ij ⎠ ·
j=1 j=1
⎛ ⎞1/2
n
(m)
⎝ |ξj − ξj |2 ⎠
j=1
1/2 ⎛ ⎞1/2 ⎛ ⎞1/2
n
(m)
n
n n
(m)
Thus |ηi − ηi |2 ≤⎝ a2ij ⎠ ⎝ |ξj − ξj |2 ⎠
i=1 i=1 j=1 j=1
n
n
Since a2ij < ∞
i=1 j=1
∞
(m) 2
||xm − x||2 −→ 0 =⇒ |ξi − ξi | −→ 0
i=1
n
(m) 2
=⇒ |ηi − ηi | −→ 0
i=1
=⇒ ||Axm − Ax||2 −→ 0 as m → ∞.
4.1.2c Example
We consider example 4.1.2a in the normed linear space C([a, b]). Since
x(s) ∈ C([a, b]) and K(t, s) is continuous in a ≤ t, s ≤ b, it follows that
y(t) ∈ C([a, b]).
Let xn , x ∈ C([a, b]) and ||xn − x|| −→ 0 as n → ∞, where ||x|| =
max |x(t)|.
a≤t≤b
Let us suppose
∞
∞
K= |aij |q < ∞, q > 1 (4.5)
i=1 j=1
∞
and x ∈ lp i.e. |ξi |p ≤ ∞.
i=1
(1) (2)
For x1 = {ξi } ∈ lp , x2 = {ξi } ∈ lp it is easy to show that A is linear.
Then, using Hölder’s inequality (1.4.3)
⎧⎛ ⎞1/q ⎛ ⎞1/p ⎫q
n n ⎪
⎨ ∞ ∞
p ⎪
⎬
|ηi |q ≤ ⎝ |aij |q ⎠ ⎝ |ξj |⎠
⎪ ⎪
i=1 i=1 ⎩ j=1 j=1 ⎭
∞
n
= ||x||q |aij |q
i=1 j=1
∞
∞
≤ ||x||q |aij |q
i=1 j=1
Linear Operators 139
1/q ⎛ ⎞1/q
∞
∞
∞
1 1
Hence ||y|| = |ηi |q ≤⎝ |aij |q ⎠ · x where + = 1.
p q
i=1 i=1 j=1
= K||xm − x||qp
Hence ||xm − x||p −→ 0 =⇒ ||Axm − Ax||q −→ 0.
Note 4.2.1. The definition 4.2.3 of a bounded linear operator is not the
same as that of an ordinary real or complex function, where a bounded
function is one whose range is a bounded set.
We would next show that a bounded linear operator and a continuous
linear operator are one and the same.
4.2.4 Theorem
In order that an additive and homogeneous operator A be continuous, it
is necessary and sufficient that it is bounded.
Proof: (Necessity) Let A be a continuous operator. Assume that it is not
bounded. Then, there is a sequence {xn } of elements, such that
Therefore, ξn −→ ξ = θ as n → 0.
On the other hand,
1
||Aξn || = ||Axn || > 1 (4.7)
n||xn ||
4.2.8 Lemma
||Ax||Ey
(i) ||A|| = sup (4.10)
x=θ ||x||Ex
(ii) ||A|| = sup AxEy (4.11)
x≤1
||Ax||Ey
sup ≤ ||A|| (4.12)
x=θ ||x||Ex
Again (ii) of lemma 4.2.7 yields
||Ax ||
> (||A|| −
)
||x ||
||Ax||Ey
Hence, sup ≥ ||A|| (4.13)
x=θ ||x||Ex
It follows from (4.12) and (4.13) that
||Ax||Ey
sup = ||A||
x=θ ||x||Ex
Next, if ||x||Ex ≤ 1, it follows from (i) of lemma 4.2.7,
sup AxEy ≤ A (4.14)
x≤1
b
Since |K(t, s)|ds is a continuous function, it attains the maximum
a
at some point t0 of the interval [a, b]. Let us take,
z0 (s) = sgn K(t0 , s)
where sgn z = z/|z|, z ∈ . +
Let xn (s) be a continuous function, such that |xn (s)| ≤ 1 and xn (s) =
z0 (s) everywhere except on a set En of measure less than 1/2M n, where
M = max |K(t, s)|. Then, |xn (s) − z0 (s)| ≤ 2 everywhere on En .
t,s
b
b
We have K(t, s)z0 (s)ds − K(t, s)xn (s)ds
a a
b
≤ |K(t, s)| |xn (s) − z0 (s)|ds
a
= |K(t, s)| |xn (s) − z0 (s)|ds
En
1 1
2 max |K(t, s)| ·=
2M n
t,s n
b b
1
Thus K(t, s)z0 (s)ds ≤ K(t, s)xn (s)ds +
a a n
1
≤ ||A|| ||xn || +
n
for t ∈ [a, b], putting t = t0
b
1
|K(t0 , s)|ds ≤ ||A|| ||xn || +
a n
Since ||xn || ≤ 1, the preceeding inequality in the limit as n → ∞ gives
rise to
b
|K(t, s)|ds ≤ ||A||,
a
b
i.e., max |K(t, s)|ds ≤ ||A|| (4.17)
t a
4.2.14 Theorem
A bounded linear operator A0 defined on a linear subset X, which
is everywhere dense in a normed linear space Ex with values in a
complete normed linear space Ey can be extended to the entire space with
preservation of norm.
Linear Operators 145
Making n → ∞ we have
But the norm of A over Ex cannot be smaller than the norm of A0 over
X, therefore we have
||A||Ex = ||A||X
The above process exhibits the completion by continuity of a bounded
linear operator from a dense subspace to the entire space.
Problems [4.1 & 4.2]
1. Let A be an nth order square matrix, i.e., A = (aij ) i=1,...n . Prove
j=1,...n
that A is linear, continuous and bounded.
2. Let B be a bounded, closed domain in 4
and let x =
T
4
(x1 , x2 , . . . , xn ) . We denote by
the space C([B]) of functions f (x)
which are continuous on B. The function φ(x) and the n-dimensional
146 A First Course in Functional Analysis
vector p(x) are fixed members of C([B]). The values of p(x) lie in B
for all x ∈ B. Show that T1 and T2 given by T1 f (x) = φ(x)f (x) and
T2 (f (x)) = f (p(x)) are linear operators.
3. Show that the matrix
⎛ ⎞
a11 a11 ··· a1n
⎜ a21 a22 ··· a2n ⎟
⎜ ⎟
A=⎜ . ⎟
⎝ .. ⎠
an1 an2 ··· ann
n
p = ∞, ||A||∞ = max |aij |
j
j=1
n
Px = x(tj )φj , x ∈ E
j=1
|f (x)| ≤ M ||x|| ∀ x ∈ E.
4.3.5 Theorem
An additive functional defined on a linear space E over 4(+) is linear
if and only if it is bounded.
The smallest of the constants M in the above inequality is called the
norm of the functional f (x) and is denoted by ||f ||.
Thus |f (x)| ≤ ||f || ||x||
Thus ||f || = sup |f (x)|
||x||≤1
|f (x)|
or in other words ||f || = sup |f (x)| = sup
||x||=1 x=θ ||x||
148 A First Course in Functional Analysis
4.3.6 Examples
4
1. Norm: The norm || · || : Ex → on a normed linear space (Ex , || · ||)
is a functional on Ex which is not linear.
2. Dot Product: Dot Product, with one factor kept fixed, defines a
4
functional f : n −→ by means of 4
n
f (x) = a · x = a i xi
i=1
|f (x)|
Therefore, sup ≤ ||a|| for x = a, f (a) = a · a = ||a||2
x=θ ||x||
f (a)
Hence, = ||a||, i.e., ||f || = ||a||.
||a||
3. Definite Integral: The definite integral is a number when we take the
integral of a single function. But when we consider the integral over a class
of functions in a function space, then the integral becomes a functional.
Let us consider the functional
b
f (x) = x(t)dt, x ∈ C([a, b])
a
b
and F (x)(s) = K(s, t)x(t)dt, x ∈ C([a, b]), s ∈ [a, b]
a
b
then |F (x)(s)| ≤ |K(s, t)| |x(t)|dt
a
b
≤ |K(s, t)| sup |x(t)|dt
a t
b b
= ||x|| |K(s, t)|dt ≤ sup |K(s, t)|dt||x||
a s a
3
|F x(s)| b
||F || = sup ≤ sup |K(s, t)|dt; s ∈ [a, b]
x∈θ ||x|| s∈[a,b] a
[Ans. ||f || = 4]
2. Let f be a bounded linear functional on a complex normed linear
space. Show that f , although bounded, is not linear (the bar denotes
the complex conjugate).
150 A First Course in Functional Analysis
3. The space C ([a, b]) is the normed linear space of all continuously
differentiable functions on J = [a, b] with norm defined by
||x|| = max |x(t)| + max |x (t)|
t∈J t∈J
= ||A|| + ||B||
Thus, the space of bounded linear operators is a normed linear space.
Linear Operators 151
4.4.2 Theorem
If Ey is complete, the space of bounded linear operators is also complete
and is consequently a Banach space.
Proof: Let us be given a Cauchy sequence {An } of linear operators.
Then with respect to the norm is a sequence of linear operators such that
||An −Am || → 0 as n, m → ∞. Hence, ||An x−Am x|| ≤ ||An −Am || ||x|| → 0,
as n, m → ∞, for any x.
Therefore the sequence {An x} of elements of Ey is a Cauchy sequence
for any fixed x. Now, since Ey is complete {An x} has some limit, y. Thus,
every y ∈ Ey is associated with some x ∈ Ex and we obtain some operator
A defined by the equation Ax = y. Such an operator A is additive and
homogeneous. Because
||An − Am || → 0 as n → ∞
4.4.4 Example
1. Let A ∈ ( 4m −→ 4n) where both 4m and 4n are normed by l1
norm.
n
Then ||A||l1 = max |aij | (4.19)
1≤j≤n
i=1
Proof: For any x ∈ 4n ,
m n
m
n
||Ax||l1 = | aij xj | ≤ |aij ||xj |
i=1 j=1 i=1 j=1
n
m
≤ |xj | |aij |
j=1 i=1
m
≤ max |aij | ||x||1
1≤j≤n
i=1
||Ax||1 n
or ||A|| = sup ≤ max |aij |.
x=θ ||x||1 1≤j≤n
i=1
Linear Operators 153
We have to next show that there exists some x ∈ n s.t. the RHS in 4
the above inequality is attained.
Let k be an index for which the maximum in (4.18) is attained; then
m
m
||Aek ||1 = |aik | = max |aij |,
1≤j≤n
i=1 i=1
⎧ ⎫
⎪
⎪ 0⎪ ⎪
⎪
⎪
⎪
⎪ 0⎪ ⎪
⎪
⎪
⎪ ⎪
⎨.⎪
⎪ . ⎬
.
where, ek = , i.e. the maximum in (4.18) is attained for the kth
⎪
⎪ 1⎪ ⎪
⎪
⎪ ⎪
.. ⎪
⎪ ⎪
⎪
⎪ .⎪⎪
⎩ ⎪
⎪ ⎭
0
coordinate vector.
Problem
1. Let A ∈ ( 4n −→ 4m ) where 4n are normed by l∞ norm.
n
Show that (i) ||A||l∞ = max |aij |
1≤i≤m
j=1
√
(ii) ||A||l2 = λ where λ is the maximum eigenvalue of AT A.
[Hint. ||Ax||2 = (Ax)T Ax = (xT AT Ax)]
2. (Limaye [33]) Let A = (aij ) be an infinite matrix with scalar entries
and
⎧ ∞
1/r
⎪
⎪
⎪
⎪ |aij |r if p = 1, 1 ≤ r < ∞
⎪
⎪ sup
⎪
⎪ j=1,2,...
⎪
⎪ i=1
⎨ ⎛ ⎞1/q
∞
αp,r = 1 1
⎪
⎪ sup ⎝ |aij |q ⎠ if 1 < p ≤ ∞, + = 1, r = ∞
⎪
⎪ p q
⎪
⎪ i=1,2,...
⎪
⎪
j=1
⎪
⎪
⎩ sup |aij | if p = 1, r = ∞
i,j=1,2,...
If αpr < ∞ then show that A defines a continuous linear map from
lp to lr and its operator norm equals αp,r .
∞
[Note that, α1,1 = α1 = sup |aij | and α∞,∞ = α∞ =
j=1,2,...
⎧ ⎫ i=1
⎨∞ ⎬
sup |aij | ].
i=1,2,... ⎩ ⎭
j=1
154 A First Course in Functional Analysis
Hence, the uniform convergence of the sequence {An } in the unit ball
||x|| < 1 of the space H does not hold.
4.5.5 Theorem
If the spaces Ex and Ey are complete, then the space of bounded linear
operators is also complete in the sense of pointwise convergence.
Proof: Let {An } of bounded linear operators converge pointwise. Since
{An } is a Cauchy sequence for every x, there exists a limit y = lim An x
n→∞
for every x.
Since Ey is complete y ∈ Ey . This asserts the existence of an operator
A such that Ax = y. That A is linear can be shown as follows
for, x1 , x2 ∈ Ex ,
A(x1 + x2 ) = lim An (x1 + x2 ) = lim (An x1 + An x2 )
n→∞ n→∞
= Ax1 + Ax2 , which shows A is additive.
Proof: Let us suppose the contrary. We show that the assumption implies
that the set {||An x||} is not bounded on any closed ball B(x0 ,
).
Any x ∈ B(x0 ,
) can be written as
x0 + ξ for any ξ ∈ Ex
||ξ||
In fact, if ||An x|| ≤ C for all n and if all x is in some ball B(x0 ,
), then
||An x0 + ξ || ≤ C
||ξ||
or, ||An ξ|| − ||An x0 || ≤ C
||ξ||
C + ||An x0 ||
or, ||An ξ|| ≤ ||ξ||
Since the norm sequence {||An x0 ||} is bounded due to {An x0 } being a
convergent sequence, it follows that
C + ||An x0 ||
||An ξ|| ≤ C1 ||ξ|| where C1 = .
The above inequality yields
||An ξ||
||An || = sup ≤ C1
ξ=0 ||ξ||
Ax = lim An x
n→∞
Making
→ 0, we get
||Ax|| ≤ M ||x||
Thus ||An x0 || ≤ k0
ξ
Now, x = x0 + r0 would belong to the B0 (x0 , r0 ) for every ξ ∈ Ex .
||ξ||
0 0
0 An ξ 0
Thus, 0 r
0 0 ||ξ|| + A 0
n 0 0 ≤ k0
x
r0
Hence, ||An ξ|| − ||An x0 || ≤ k0
||ξ||
k0 + ||An x0 || 2k0
or ||An ξ|| ≤ ||ξ|| ≤ ||ξ||
r0 r0
||An ξ|| 2k0
Thus, ||An || = sup ≤
ξ=0 ||ξ|| r0
Hence, {||An ||} is bounded.
158 A First Course in Functional Analysis
4.5.9 Remark
The theorem does not hold unless Ex is a Banach space, as it follows
from the following example.
4.5.10 Examples
1. Let Ex = {x = {xj } : only a finite number of xj ’s are non-zero}.
||x|| = sup |xj |
j
Consider the mapping, which is a linear functional defined by fn (x) =
n
xj then
j=1
n n m
|fn (x)| = xj ≤ |xj | ≤ |xj |, since xi = 0 for i > m.
j=1 j=1 j=1
m
|fn (x)| ≤ |xj | ≤ m||x||
j=1
Hence ||fn || ≤ n
Next, consider the element ξ = {ξ1 , ξ2 , . . . , ξi , . . .}
where ξi = 1 1 ≤ i ≤ n
= 0 i > n.
||ξ|| = 1
f (ξi ) = ξi = n = n||ξ||
n
|fn (ξ)|
=n
||ξ||
Thus {fn (x)} is bounded but {||fn ||} is not bounded. This is because
Ex = {x = {xj } : only for a finite number of xj ’s is non-zero} is not a
Banach space.
∞
2. Let Ex be the set of polynomials x = x(t) = pn tn where pn = 0 for
n=0
n > Nx .
Let ||x|| = max[|pn |, n = 1, 2, . . .]
Linear Operators 159
n−1
Let fn (x) = pk . The functionals fn are continuous linear functionals
k=0
on Ex . Moreover, for every x = p0 + p1 t + · · · + pm tm , it is clear that for
every n, |fn (x)| ≤ (m + 1)||x||, so that {|fn (x)|} is bounded. For fn (x) we
choose x(t) = 1 + t + · · · + tn .
|fn (x)|
Now, ||fn || = sup ≥ n since ||x|| = 1 and |fn (x)| = n.
x=θ ||x||
Hence {||fn ||} is unbounded.
n
(n)
Ln f = wk (t)f (tnk ) (4.20)
k=1
(n) pn (t)
where wk (t) = (4.21)
pn (t)(t − tnk )
E
n
and pn (t) = (t − tnk ). (4.22)
k=1
4.6.2 Theorem
We are given some points on the segment [0, 1] forming the infinite
triangular matrix,
⎡ ⎤
t11 0 0 ··· 0
⎢ 2 ⎥
⎢ t1 t22 0 ··· 0⎥
⎢ ⎥
T =⎢
⎢ t1
3
t32 t33 ··· 0⎥ ⎥. (4.23)
⎢ ⎥
⎣· · · ··· ··· ··· · · ·⎦
··· ··· ··· ··· ···
160 A First Course in Functional Analysis
For a given function, f (t) defined on [0, 1], we construct the Lagrangian
interpolation polynomial Ln f whose partition points are the points of nth
row of (4.23),
n
(n)
Ln f = wk (t)f (tnk )
k=1
(n) pn (t) E
n
where wk = , pn (t) = (t − tnk ).
pn (t)(t − tnk )
k=1
For every choice of the matrix (4.23), there is a continuous function f (t)
s.t. Ln f does not uniformly converge to f (t) as n → ∞.
Proof: Let us consider Ln as an operator mapping the function f (t) ∈
C([0, 1]) into the elements of the same space and put
n
(n)
λn = max λn (t) where λn (t) = |wk (t)|
t
k=1
n
(n) n
Now, ||Ln f || = max wk (t)f (tk )
t
k=1
n
(n)
≤ max |wk (t)| max |f (tnk )|
t t
k=1
n
(n)
= λn ||f ||, where λn = max |wk (t)|
t
k=1
On C([0, 1]) ||f || = max |f (t)|
t
||Ln f ||
||Ln || = sup ≤ λn
f =θ ||f ||
taken as ,
eint
en (t) = √ , n = 0, 1, 2, . . .
2π
If x(t) ∈ L2 ([−π, π]) then x(t) can be written as
∞
xn en
n=−∞
π
1
where xn = x, en = √ x(t)e−int dt
2π −π
π π
1 1
=√ x(t) cos ntdt − √ x(t) sin ntdt
2π −π 2π −π
= cn − idn (say)
Then, xn = cn + idn
∞ ∞ ∞
2cn cos nt 2dn sin nt
Thus xn en = x0 e0 + √ + √ (4.24)
n=−∞ n=1
2π n=1
2π
∞
1
x(t) = √ xn eint (4.25)
2π n=−∞
1 n
un (t) = √ xk eikt (4.26)
2π k=−∞
∞ ∞
2cn cos nt 2dn sin nt
x(t) = x0 + √ + √ (4.27)
n=1
2π n=1
2π
π
1
where x0 = √ x(t)dt (4.28)
2π −π
π
cn = x(t) cos ntdt (4.29)
−π
π
dn = x(t) sin ntdt (4.30)
−π
For x(t) ∈ E
||x|| = max |x(t)| (4.31)
−π≤t≤π
1
n n
1
Now, 2 sin t cos mt = 2 sin t cos mt
2 m=1 m=1
2
n
1 1
= sin m + t − sin m − t
m=1
2 2
1 1
= − sin t + sin n + t
2 2
It may be noted that, except for the end terms, all other intermediate
terms in the summation vanish in pairs.
Dividing both sides by sin 12 t and adding 1 to both sides, we have
! "
n
sin n + 12 t
1+2 cos mt =
m=1
sin 12 t
We would next show that un is bounded. It follows from (4.29) and the
above integral,
π
1
|un (x)| ≤ √ |x(t)| |Dn (t)|dt
2π −π
π
1
≤√ |Dn (t)|dt||x||
2π −π
π
|un (x)| 1
Therefore, ||un || = sup ≤√ |Dn (t)|dt
x=θ ||x|| 2π −π
1
= √ ||Dn (t)||1
2π
where || · ||1 denotes L1 -norm.
Actually the equality sign holds, as we shall prove.
Let us write |Dn (t)| = y(t)Dn (t) where y(t) = +1 at every t at which
Dn (t) ≥ 0 and y(t) = −1 otherwise y(t) is not continuous, but for any given
> 0 it may be modified to a continuous x of norm 1, such that for this x
we have
π
un (x) − √1
1 π
|Dn (t)|dt = √ (x(t) − y(t))Dn (t)dt <
2π −π 2π −π
If follows that π
1
||un || = √ |Dn (t)|dt
2π −π
We next show that the sequence {||un ||} is unbounded. Since sin u ≤ u,
0 ≤ u ≤ π, we note that
3
π
! 1
" 3
π/2
sin n + 2 t sin(2n + 1)u
dt ≥ 4 du
−π sin 12 t 0 u
(2n+1)π
2 | sin v|
=4 dv
0 v
2n (k+1) π
2 | sin v|
=4 dv
kπ
2
v
k=0
2n
(k+1)π
1 2
≥4 | sin v|dv
(k + 1) π2 kπ
2
k=0
2n
8 1
= −→ ∞ as n → ∞
π (k + 1)
k=0
∞
1
Because the harmonic series diverges.
k
k=1
164 A First Course in Functional Analysis
Hence {||un ||} is unbounded. Since E is complete this implies that there
exists no c > 0 and finite such that ||un (x)|| ≤ c holds for all x. Hence
there must be an x ∈ E such that {||un (x )||} is unbounded. This implies
that the Fourier series of that x diverges at t = 0.
Problems [4.5 & 4.6]
1. Let Ex be a Banach space and Ey a normed linear space. If {Tn } is
a sequence in (Ex → Ey ) such that T x = limn→∞ Tn x exists for each
x in Ex , prove that T is a continuous linear operator.
2. Let Ex and Ey be normed linear spaces and A : Ex → Ey be a linear
N
operator with the property that the set {||Axn || : n ∈ } is bounded
whenever xn → θ in Ex . Prove that A ∈ (Ex → Ey ).
3. Given that Ex and Ey are Banach spaces and A : Ex → Ey is a
bounded linear operator, show that either A(Ex ) = Ey or is a set of
the first category in Ey .
[Hint: A set X ⊆ Ex is said to be of the first category in Ex if it is
the union of countably many nowhere dense sets in Ex ].
4. If Ex is a Banach space, and {fn (x)} a sequence of continuous linear
functionals on Ex such that {|fn (x)|} is bounded for every x ∈ Ex ,
then show that the sequence {||fn ||} is bounded.
[Hint: Consult theorem 4.5.7]
5. Let Ex and Ey be normed linear spaces. E be a bounded, complete
convex subset of Ex . A mapping A from Ex to Ey is called affine if
4.7.8 Lemma
If A has a left inverse B and a right inverse C, then B = C.
For B = B(AC) = (BA)C = C.
If A has a left inverse as well as a right inverse, then A is said to have
an inverse and the inverse operator is denoted by A−1 .
Thus if A−1 exists, then by definition A−1 A = AA−1 = I.
4.7.9 Inverse operators and algebraic equations
Let Ex and Ey be two Banach spaces and A be an operator s.t.
A ∈ (Ex −→ Ey ).
We want to know when one can solve
Ax = y (4.33)
m, some constant. Then the inverse bounded linear operator A−1 exists.
Proof: The condition (4.34) implies that A maps Ex onto Ey in a one-
to-one fashion. If Ax1 = y and Ax2 = y, then A(x1 − x2 ) = θ yields
Linear Operators 167
4.7.12 Theorem
Let a bounded linear operator A map E into E and let ||A|| ≤ q < 1.
Then the operator I +A has an inverse, which is a bounded linear operator.
Proof: In the space of operators with domain E and range as well in E,
we consider the series
I − A + A2 − A3 + · · · + (−1)n An + · · · (4.35)
= lim (I − An+1 ) = I
n→∞
−1
Hence S = (I + A)
Let x1 = (I + A)−1 y1 , x2 = (I + A)−1 y2 , x1 , x2 , y1 , y2 ∈ E
Then y1 + y2 = (I + A)(x1 + x2 )
or, x1 + x2 = (I + A)−1 y1 + (I + A)−1 y2 = (I + A)−1 (y1 + y2 ).
Hence, S is a linear operator. Moreover,
∞ ∞
1
||S|| ≤ ||An || ≤ qn = , 0<q<1 (4.36)
n=0 n=0
1−q
Proof: Since ||I − A−1 C|| = ||A−1 (A − C)|| ≤ αβ < 1 and A−1 C =
I − (I − A−1 C), it follows from theorem 4.7.12 that A−1 C has a bounded
inverse and hence C has a bounded inverse.
4.7.14 Example
Consider the integral operator
1
Cx = x(s) − K(s, t)x(t)dt (4.36 )
0
Linear Operators 169
with continuous kernel K(s, t) which maps the space C([0, 1]) into C([0, 1]).
Let K0 (s, t) be a degenerate kernel close to K(s, t). That is, K0 (s, t) is
n
of the form ai (s)bi (t). In such a case, the equation
i=1
1
Ax = x(s) − K0 (s, t)x(t)dt = y (4.37)
0
k
Making k −→ ∞, we have, y = lim yi .
k→∞
i=1
Let xk = A−1 yk , then
nl
||xk ||Ex = ||A−1 yk ||Ex ≤ n||yk ||Ey ≤
2k−1
k
Expressing sk = xi
i=1
k+p
nl 1 nl
||sk+p − sk || = || xi || ≤ 1 − p < k−1
2k−1 2 2
i=k+1
k
Ax = A lim xi = lim Axi
k→∞ k→∞
i=1 i=1
k
= lim yi = y
k→∞
i=1
k
k
Hence, ||A−1 y|| = ||x|| = lim || xi || ≤ lim ||xi ||
k k
i=1 i=1
Linear Operators 171
∞
nl
≤ = 2nl = 2n||y||
i=1
2i−1
4.7.16 Examples
t
1. Let E = C([0, 1]) and let Au = u(ζ)dζ.
0
Thus, A is a bounded linear operator mapping C([0, 1]) into C([0, 1]).
d
A−1 given by A−1 u = u(t)
dt
is an unbounded operator defined on the linear subspace of continuously
differentiable functions such that u(0) = 0.
2. Let E = C([0, 1]). The operator B is given by
Bu = f (x) (4.41)
d2
where B = −
dx2
We change the order of integration and use Fubini’s theorem [see 10.5].
This leads to
x x
u(x) = − f (t)dt ds + C1 x
0 t
172 A First Course in Functional Analysis
x
=− (x − t)f (t)dt + C1 x (4.42)
0
1
u(1) = 0 ⇒ C1 = (1 − t)f (t)dt (4.43)
0
Using (4.43), (4.42) reduces to
x 1
u(x) = t(1 − x)f (t)dt + x(1 − t)f (t)dt
0 x
1
= K(x, t)f (t)dt = B −1 f (4.44)
0
K(x, t), the kernel of the integral equation (4.44) is given by
3
t(1 − x) 0 ≤ t ≤ x
K(x, t) =
x(1 − t) x ≤ t ≤ 1
We note that B is not a bounded operator. For example, take
u(x) = sin nπx, u(x) ∈ D(B)
d2
Now, 2
sin nπx = −n2 π 2 sin nπx.
dx
0 2 0
0 d 0
Then, 0 0 dx2 (sin nπx) 0 = n2 π 2 −→ ∞ as n → ∞.
0
On the other hand
1
−1 1
||B f || ≤ |K(x, t)|dt||f || ≤ ||f ||
0 8
−1
Hence B is bounded.
4.7.17 Operators depending on a parameter
Let us consider the equation of the form
Ax − λx = y or, (A − λI)x = y (4.45)
solution x = θ. Those λ for which equation (4.45) has a unique solution for
every y and the operator Rλ is bounded, are called regular values.
4.7.20 Definition: eigenvector, eigenvalue or characteristic
value, spectrum
If the homogeneous equation (4.46) or the equation Ax = λx has a non-
trivial solution x, then that x is called the eigenvector. The values of
λ corresponding to the non-trivial solution, the eigenvector, are called the
eigenvalues or characteristic values.
The collection of all non-regular values of λ is called the spectrum of
the operator A.
4.7.21 Theorem
If in the equation (A − λI)x = y, (1/|λ|)||A||q < 1 holds for λ, thus
A − λI has an inverse operator; moreover,
1 A A2
Rλ = − 1 + + 2 + ···
λ λ λ
If λ is a regular value, then λ + Δλ for |Δλ| < ||(A − λI)−1 ||−1 is also a
regular value. This implies that the collection of regular values is an open
set and hence the spectrum of an operator is a closed set.
Proof: The equation (A − λI)x = y can be written as
A 1
− I− x= y
λ λ
1
Thus, by theorem 4.7.12 we can say, if |λ| ||A|| = q < 1, then (I − A/λ)
will have an inverse and the concerned values of λ will be its regular values.
If in theorem 4.7.13 we take C = A−(λ+Δλ)I then if ||C −(A−λI)|| =
|Δλ| < ||(A − λI)−1 ||−1 then we conclude that C = A − (λ + Δλ)I has an
inverse and (λ + Δλ) is also a regular value.
4.7.22 Example
Let us consider the Fredholm integral equation of the second kind:
b
φ(x) − λ K(x, s)φ(s)ds = f (x) (4.47)
a
b
If we denote Aφ by K(x, s)φ(s)ds, equation (4.47) can be written as
a
(I − λA)φ = f (4.48)
The solution of the equation (4.47) can be expressed in the form of an
infinite series
∞ b
φ(x) = f (x) + λm Km (x, s)f (x)ds (see Mikhlin [37]) (4.49)
m=1 a
174 A First Course in Functional Analysis
Rλ f = (I − λA)−1 f = [I + λA + λ2 A2 + · · · + λp Ap + · · · ]f
b
where Ap f = Kp (t, s)f (s)ds is the pth iterate of the kernel K(t, s).
a
1
Thus if ||A|| < , we get
|λ|
b
φ(x) = f (x) + λ R(x, s, λ)f (s)ds
a
Here, R(x, s, λ), the resolvent of the kernel K(t, s) is defined by,
Problems
1. Let Ex and Ey be normed linear spaces over 4(+). A : Ex −→ Ey is
a given linear operator. Show that
4
2. L( n ) denotes the space of linear operators mapping n → n . 4 4
Suppose that the mapping A : D ⊂ Rm → L(Rn ) is continuous
at a point x0 ∈ D for which A(x0 ) has an inverse. Then, show that
there is a δ > 0 and a ν > 0 so that A(x) has an inverse and that
||A(x)−1 || ≤ ν ∀ x ∈ D ∩ B(x0 , δ)
5. Use the result of problem 4 to find the solution of the linear differential
equation
dU
− λU = f, u(t) ∈ C 1 [0, 1], f ∈ C([0, 1]) and |λ| < 1
dt
6. Let m be the space of bounded number sequences, i.e., for x ∈ m =⇒
x = {ξi }, |ξi | ≤ Kx , ||x|| = sup |ξi |.
i
In m the shift operator E is defined by
Ex = (0, ξ1 , ξ2 , ξ3 , . . .) for x = (ξ1 , ξ2 , ξ3 , . . .)
The two sequences {ei } and sequence of functionals {fi } such that,
(4.52) is true, are called biorthogonal sequences.
4.8.3 Lemma
For every functional f defined on the Banach space E, we can find
coefficients ci = f (ei ), where {ei } is a Schauder basis in E, such that
∞
∞
f= f (ei )fi = ci fi ,
i=1 i=1
Writing, f (ei ) = ci ,
∞
∞
f (x) = ci fi (x) or, f = ci fi (4.53)
i=1 i=1
t, (1 − t), u00 (t), u10 (t), u11 (t), u20 (t), u21 (t), . . . , u22 (t) (4.54)
a01
a0 a1
O t t
1/2 3/4 1 O 1/4 1/2 3/4 1
Fig. 4.2 Fig. 4.3
x1(t )
}
1
t
0 1
x2(t )
}
1
0 t
1
n = 1 x (t )
}
3
j=1 √2
t
x4(t ) n = 1
t j=2
Fig. 4.4
CHAPTER 5
LINEAR
FUNCTIONALS
179
180 A First Course in Functional Analysis
x1 + t1 x0 = x2 + t1 x0 =⇒ x1 = x2
x1 − x 2
Now, x1 + t1 x0 = x2 + t2 x0 or, x0 =
t2 − t 1
showing that x0 ∈ L since x1 , x2 ∈ L. Hence x1 = x2 and t1 = t2 , i.e., the
representation of u is unique. Let us take two elements, x and x ∈ L.
sup{f (x) − ||f || ||x + x0 ||} ≤ inf {f (x) + ||f || ||x + x0 ||}
x∈L x∈L
sup{f (x) + ||f || ||x + x0 ||} ≤ c ≤ inf {f (x) + ||f || ||x + x0 ||} (5.1)
x∈L x∈L
Thus, f0 (u) is additive. To show that f0 (u) is bounded and has the
same norm as that of f (x) we consider two cases:
f (x) ≤ r||f ||
5.1.8 Example
|| · || is a sublinear functional.
F (x) = f (x), ∀ x ∈ E
g1 (y + αy1 ) = F (y) + αc
Hence, f is complex-linear.
5.2.2 Hahn-Banach theorem (generalized)
Let E be a real or complex vector space and p be a real-valued functional
on E which is subadditive, i.e., for all x, y ∈ E
p(x + y) ≤ p(x) + p(y) (5.11)
and for every scalar α satisfies,
p(αx) = |α|p(x) (5.12)
Proof: (a) Real Vector Space: Let E be real. Then (5.13) yields f (x) ≤
p(x) ∀ x ∈ L. It follows from theorem 5.1.9 that f can be extended to a
linear functional F from L to E such that
F (x) ≤ p(x) ∀ x ∈ E (5.15)
Linear Functionals 187
F (x) = |F (x)|eiθ ,
Furthermore, p(x + y) = ||f ||L ||x + y|| ≤ ||f ||L (||x|| + ||y||)
= p(x) + p(y), ∀ x, y ∈ E
p(αx) = ||f ||L ||αx|| = |α| ||f ||L ||x||)
= |α|p(x) ∀ x ∈ E
Thus conditions (5.11) and (5.12) of theorem 5.2.2 are satisfied. Hence,
the above theorem can be applied, and we get a linear functional F on E
which is an extension of f and satisfies.
Linear Functionals 189
5.2.13 Remark
(i) If X is an absorbing set, μX (x) < ∞ for every x ∈ E.
(ii) If X is an absorbing set then, θ ∈ X and
(iii) If X is a normed linear space, then every open set containing θ is an
absorbing set.
Proof: We prove (iii). Since an open set containing ‘θ’ contains an open
ball B(θ,
), if x ∈ B(θ,
) then ||x|| <
. Therefore, for any α > 1
||α−1 x|| ≤ <
α
x = p0 (p−1
0 x) + (1 − p0 ) · θ ∈ X
B(x1 − x2 , ) = B(x1 , ) − x2 ⊂ X1 − x2
Thus, X1 − X2 is open.
*
Hence, X1 − X2 = (x1 − x2 ) is open in E.
x1 ∈X1
x2 ∈X2
Also, θ ∈
/ X1 − X2 , since X1 ∩ X2 = Φ.
Let X = X1 − X2 + u0 where u0 = x2 − x1 . Then X is an open convex
set with θ ∈ X. Hence, X is an absorbing set as well. Let μX be the
Minkowski functional of X.
In order to obtain the required functional, we apply the theorem 5.1.9.
4
Let E0 = span {u0 }, p = μX and the linear functional f0 : E −→ defined
by
f0 (λu0 ) = λ, λ ∈ 4 (5.26)
Since X1 ∩ X2 = Φ and u0 = x2 − x1 ∈
/ X, by lemma 5.2.14, we have
μX (u0 ) ≥ 1 and hence,
|f (x)| ≤ 1 ∀ x ∈ X ∩ (−X)
Prove that L = E.
7. Let E be a normed linear space. For every subspace L of E and every
functional f defined on E, prove that there is a unique Hahn-Banach
extension of f to E if and only if E ∗ is strictly convex that is, for
f1 = f2 in E, with ||f1 || = 1 = ||f2 || we have ||f1 + f2 || < 2.
[Hint: If F1 and F2 are extensions of f , show that F1 + F2 /2 is also
a continuous linear extension of f and the strict convexity condition
is violated].
The supremum being taken over all partitions 5.28 of the interval [a, b].
If Var (w) < ∞ holds, then w is said to be a function of bounded
variation.
All functions of bounded variation on [a, b] form a normed linear space.
A norm on this space is given by
The normed linear space thus defined is denoted by BV ([a, b]), where
BV suggests ‘bounded variation’.
We now obtain the concept of a Riemann-Stieljes integral as follows.
Let x ∈ C([a, b]) and w ∈ BV ([a, b]). Let Pn be any partition of [a, b] given
by (5.28) and denote by η(Pn ) the length of a largest interval [tj−1 , tj ] that
is,
η(Pn ) = max(t1 − t0 , t2 − t1 , . . . , tn − tn−1 ).
For every partition Pn of [a, b], we consider the sum,
n
S(Pn ) = x(tj )[w(tj ) − w(tj−1 )] (5.31)
j=1
There exists a number I with the property that for every
> 0 there is
a δ > 0 such that
η(Pn ) < δ =⇒ |I − S(Pn )| <
I is called the Riemann-Stieljes integral of x over [a, b] with respect to
w and is denoted by
b
x(t)dw(t) (5.32)
a
196 A First Course in Functional Analysis
Thus, we obtain (5.32) as the limit of the sum (5.31) for a sequence
{Pn } of partitions of [a, b] satisfying η(Pn ) → 0 as n → ∞.
In case w(t) = t, (5.31) reduces to the familiar Riemann integral of x
over [a, b].
Also, if x is continuous on [a, b] and w has a derivative which is integrable
on [a, b] then
b b
x(t)dw(t) = x(t)w (t)dt (5.33)
a a
We show that the integral (5.32) depends linearly on x, i.e., given
x1 , x2 ∈ C([a, b]),
b b b
[px1 (t) + qx2 (t)]dw(t) = p x1 (t)dw(t) + q x2 (t)dw(t)
a a a
where p, q ∈ 4
The integral also depends linearly on w ∈ BV ([a, b]) because for all
w1 , w2 ∈ BV ([a, b]) and scalars r, s
b b b
x(t)d(rw1 + sw2 )(t) = r x(t)dw1 (t) + s x(t)dw2 (t)
a a a
5.3.2 Lemma
For x(t) ∈ C([a, b]) and w(t) ∈ BV ([a, b]),
b
x(t)dw(t) ≤ max |x(t)| Var (w) (5.34)
a t∈[a,b]
n
= x(t0 )w(t1 ) + x(tj−1 )[w(tj ) − w(tj−1 )]
j=2
n
= x(tj−1 )[w(tj ) − w(tj−1 )] (5.42)
j=1
Linear Functionals 199
where the last equality follows from w(t0 ) = w(a) = 0. We now choose any
sequence {Pn } of partitions of [a, b] such that η(Pn ) → 0 (It is to be kept in
mind that tj depends on the particular partition Pn ). As n → ∞ the sum
on the right-hand side of (5.42) tends to the integral in (5.26) and (5.26)
follows provided F (zn ) −→ F (x), which equals f (x) since x ∈ C([a, b]).
We need to prove that F (zn ) −→ F (x). Keeping in mind the definition
of ut (ξ) (fig. (5.1)), we note that (5.41) yields zn (a) = x(a) · 1 since the sum
in (5.41) is zero at t = a. Hence zn (a) − x(a) = 0. Moreover by (5.41) if
tj−1 ≤ ξ < tj , then we obtain zn (t) = x(tj−1 ) · 1. It follows that for those
t,
|zn (t) − x(t)| = |x(tj−1 ) − x(t)|
|f (x)|
||f || = sup ≤ Var (w) (5.43)
x=θ ||x||
Note 5.3.1. We note that w in the theorem is not unique. Let us impose
on w the following conditions
(i) w is zero at a and continuous from the right
n
f (x) = f ξi ei = ξf (ei ) (5.44)
i=1 i=1
n
= ξ i fi where fi = f (ei )
i=1
n
|φ(x)| = ξi φi ≤ |ξi ||φi | ≤ |φi | ||x||,
i=1 i=1 i=1
Linear Functionals 201
|φ(x)|
n
Hence, ||φ|| = sup ≤ |φi | (5.45)
x=θ ||x|| i=1
n
On the other hand, if we select an element x0 = sgn φi ei ∈ 4n, then
i=1
||x0 || = 1 and
n
n
φ(x0 ) = sgn φi · φi (ei ) = sgn φi · φi
i=1 i=1
n
= |φi |||x0 ||
i=1
n
Hence, ||φ|| ≥ |φi | (5.46)
i=1
From (5.45) and (5.46), it follows that
n
||φ|| = |φi |
i=1
Since this series must converge for every number sequence {ξk }, the uk
must be equal to zero from a certain index onwards and consequently
m
f (x) = ξk uk
k=1
202 A First Course in Functional Analysis
m
m
Moreover, φ(λx) = λξk uk = λ ξk uk = λφ(x).
k=1 k=1
Hence, φ(x) is a linear functional on s.
It therefore follows that every linear functional defined on s has the
general form given by (5.47) where m and uk , k = 1, 2, . . . , m are uniquely
defined by (5.47).
5.4.3 The general form of linear functionals on lp
Let f (x) be a bounded linear functional defined on lp . Since the elements
ei = {ξij } where ξij = 1 for i = j and ξij = 0 for i = j, form basis of lp ,
every element x ∈ lp can be written in the form
∞
x= ξi ei
i=1
(n)
Let us put xn = {ξi }, where
3
(n)
|ui |q−1 sgn ui , if i ≤ n
ξi =
0 if i > n
q is chosen such that the equality [(1/p) + (1/q)] = 1 holds
n
(n)
n
n
f (xn ) = ui ξi = |ui |q−1 ui sgn ui = |ui |q (5.49)
i=1 i=1 i=1
∞
1/p
(n)
f (xn ) ≤ ||f || ||xn || = ||f || |ξi |p
i=1
1/p
1/p
n
n
= ||f || |ui | p(q−1)
= ||f || |ui | q
i=1 i=1
n
1/p
n
Thus |ui |q ≤ ||f || |ui |q
i=1 i=1
1 1
Since + = 1
p q
n
1/q
Thus, |ui |q ≤ ||f ||.
i=1
i=1
204 A First Course in Functional Analysis
∞
1/q
Consequently, ||f || ≤ |ui |q (5.52)
i=1
It follows from (5.50) and (5.52)
∞
1/q
||f || = |ui | q
i=1
5.4.4 Corollary
Every linear functional defined on l2 can be written in the general form
∞
f (x) = ui ξ i
i=1
∞
∞
1/2
2 2
where |ui | < ∞ and ||f || = |ui |
i=1 i=1
x1
If = u, then
||x1 ||2
f (x) = x, u, (5.53)
i.e., we get the representation of an arbitrary functional as an inner product
of the element x and a fixed element u. The element u is defined uniquely
by f because if f (x) = x, v, then x, u − v = 0 for every x ∈ H, implying
u = v.
Further (5.53) yields,
|f (x)|
sup ≤ ||u|| or ||f || ≤ ||u|| (5.54)
x=θ ||x||
Since, on the other hand, f (u) = u, u = ||u||2 , it follows that ||f ||
cannot be smaller than ||u||, hence ||f || = ||u||.
Thus, every linear functional f (x) in a Hilbert space H can be
represented uniquely in the form f (x) = x, u, where the element u is
uniquely defined by the functional f . Moreover, ||f || = ||u||.
Problem
∞
1. If l1 is the space of real elements x = {ξi } where |ξi | < ∞, show
i=1
that a linear functional f on l1 can be represented in the form
∞
f (x) = ck ξk
k=1
This shows that the norm of f is the Euclidean norm and ||f || = ||a||
where a = {ai } ∈ . 4
Linear Functionals 207
4 4
Hence, the mapping of n onto n defined by f −→ a = {ai } where
ai = f (ei ), is norm preserving and, since it is linear and bijective, it is an
isomorphism.
5.6.2 Space l1 : the dual space of l1 is l∞
Let us take a Schauder basis {ei } for l1 , where ei = (δij ), δij stands for
the Kronecker δ-symbol.
Thus every x ∈ l1 has a unique representation of the form
∞
x= ξi ei (5.57)
i=1
For any bounded linear functional f defined on l1 i.e. for every f ∈ l1∗
we have
∞
∞
f (x) = ξi f (ei ) = ξi ai (5.58)
i=1 i=1
If y = {ηi } ∈ l1 , then
∞
∞
∞
φ(x + y) = (ξi + ηi )φ(ei ) = ξ i bi + η i bi
i=1 i=1 i=1
= φ(x) + φ(y) showing φ is additive
|φ(x)|
Therefore, ||φ|| = sup ≤ sup |bi | < ∞ since b = {bi } ∈ l∞ . Thus
x=θ ||x|| i
∗
φ is bounded linear and φ ∈ l1 .
We finally show that the norm of f is the norm on the set l∞ . From
(5.58), we have,
∞
∞
|f (x)| = ξi ai ≤ sup |ai | |ξi | = ||x|| sup |ai |
i i
i=1 i=1
f (x)
Hence, ||f || = sup ≤ sup |ai |.
x=θ ||x|| i
It follows from (5.59) and this above inequality,
||f || = sup |ai |,
i
We consider any f ∈ where lp∗ lp2 is the conjugate (or dual) space of lp .
Since f is linear and bounded,
∞
∞
f (x) = ξi f (ei ) = ξi ai (5.61)
i=1 i=1
where ai = f (ei ).
1 1
Let q be the conjugate of p i.e. + = 2.
p q
(n)
Let xn = {ξ i } with
3
(n) |ai |q /ai if i ≤ n and ai = 0
ξi = (5.62)
0 if i > n or ai = 0
∞
n
(n)
Then f (xn ) = ξ i ai = |ai |q
i=1 i=1
Using (5.62) and that (q − 1)p = q, it follows from the above,
Linear Functionals 209
∞
1/p
(n) p
f (xn ) ≤ ||f || ||xn || = ||f || |ξ |
i=1
1/p
n
= ||f || |ai | p(q−1)
i=1
1/p
n
= ||f || |ai |q
i=1
1/p
n
n
Hence, f (xn ) = |ai | ≤ ||f ||
q
|ai | q
i=1 i=1
∞
1/q
= |bi | q
||x||
i=1
Then we have two cases: namely (i) g(t) is fixed and x(t) varying or
(ii) x(t) is fixed and g(t) varies. Now, since f (x) ∈ , f (x) can be treated 4
Linear Functionals 211
f
x Fx
E E* E**
Fig. 5.2
4
Hence, n is reflexive.
Note 5.6.3 Every finite dimensional normed linear space is reflexive. We
know that in a finite dimensional normed linear space E, every linear
functional on E is bounded, so that the reflexivity of E follows.
2. The space lp (p > 1)
1 1
In 5.6.3, we have seen that lp∗ = lq , + = 1
p q
∗∗ ∗ ∗ ∗
Therefore, lP = (lp ) = (lq ) = lp
Hence, lp is reflexive.
3. The space C([0, 1]) is non-reflexive. For that see 6.2.
5.6.10 Theorem
A normed linear space is isometrically isomorphic to a dense subspace
of a Banach space.
Proof: Let E be a normed linear space. If φ : E → E ∗∗ be the natural
embedding, then E and φ(E) are isometrically isomorphic spaces. But
φ(E) is a dense subspace of φ(E) and φ(E) is a closed subspace of the
Banach space E ∗∗ , it follows that φ(E) itself is a Banach space. Hence E
is isometrically isometric to the dense subspace φ(E) of the Banach space
φ(E).
We next discuss the relationship between separability and reflexivity of
a normed linear space.
5.6.11 Theorem
Let E be a normed linear space and E ∗ be its dual. Then E ∗ is separable
⇒ E is separable.
N
Proof: Since E ∗ is separable, ∃ a countable set S = {fn : fn ∈ E ∗ , n ∈ }
such that S is dense in E ∗ , i.e., S = E ∗ .
214 A First Course in Functional Analysis
f = A*g g
( )
Fig. 5.3
5.6.16 Examples
4 4
1. Let A be an operator in ( n → n ), where n is an n-dimensional 4
space. Then A is defined by a matrix (aij ) of order n and equality y = Ax
where x = {ξ1 , ξ2 , . . . , ξn } and y = {η1 , η2 , . . . , ηn } such that
n
ηi = aij ξj
j=1
Consider a functional f ∈ n∗
(=4 4n ) since 4n is self-conjugate;
n
f = (f1 , f2 , . . . , fn ), f (x) = fi ξi .
i=1
n
n
n
Hence, f (Ax) = fi ηi = fi aij · ξj
i=1 i=1 j=1
n
n
n
n
= aij fi ξj = aij fi ξj
i=1 j=1 j=1 i=1
n
= φ j ξj (5.72)
j=1
216 A First Course in Functional Analysis
n
where φj = aij fi (5.73)
i=1
= T ∗ v, f .
1
where, T ∗ v(s) = K(t, s)v(t)dt
0
Thus, in the given case, the adjoint operator is also an integral operator,
the kernel K(t, s) which is obtained by interchanging the arguments of
K(s, t). K(t, s) is called the transpose of the kernel K(s, t).
5.6.17 Theorem
Given A, a bounded linear operator mapping a normed linear space Ex
into a normed linear space Ey , its adjoint A∗ is also a bounded linear
operator, and ||A|| = ||A∗ ||.
Let f1 = A∗ g1 and f2 = A∗ g2 .
Hence, g1 (y) = g1 (Ax) = f1 (x), x ∈ Ex , f1 ∈ Ex∗ , y ∈ Ey , g1 ∈ Ey∗ .
Also g2 (y) = g2 (Ax) = f2 (x).
Now, g1 and g2 are linear functionals and hence f1 and f2 are linear
functionals.
Linear Functionals 217
g(Ax) = f0 (x), x ∈ Lx
Let x1 , x2 ∈ Lx .
n
n
y = Ax = lim αi Aei = lim αi pki ek (5.78)
n n
i=1 i=1 k=1
∞
Thus, y = βk ek (5.79)
k=1
∞
where, βk = pki αi (5.80)
i=1
3
1 if j = k
φj (ek ) = (5.81)
0 if j = k
Then (5.79) and (5.80) imply,
3 ∞
n
βm = φm (y) = φm lim αi pki ek
n
i=1 k=1
3 ∞
n
= lim φm αi pki ek
n
i=1 k=1
n
n
= lim αi pki φm (ek )
n
i=1 k=1
n
= lim pmi αi (5.82)
n
i=1
n
= lim pki αi ck = lim pki ck αi
n→∞ n→∞
k=1 i=1 i=1 k=1
Consequently, di αi = lim pki ck αi (5.83)
n→∞
i=1 i=1 k=1
220 A First Course in Functional Analysis
SPACE OF BOUNDED
LINEAR
FUNCTIONALS
|f (x)|
||f || = sup , x ∈ E.
x=θ ||x||
221
222 A First Course in Functional Analysis
|f (x)|
||x|| = sup , x ∈ E.
f =θ ||f ||
(ii) The dual of c00 with the norm || · ||p is linearly isometric to lq .
(iii) The dual of c0 with the norm || ||∞ is linearly isometric to l1 .
∞
n
Proof: (i) If we replace the summation ξi yi with the summation ξ i yi
i=1 i=1
in theorem 6.1.3 and follow its argument we get the result.
(ii) If 1 ≤ p < ∞. Then c00 is a dense subspace of lp , so that the dual
of c00 is linearly isometric to lq by theorems 6.1.2(a) and 6.1.3.
Let p = ∞, so that q = 1. Consider y ∈ l1 and define,
∞
fy (x) = xj yj , x ∈ c00 .
j=1
Following 6.1.3 we show that fy ∈ (c00 )∗ and ||fy || ≤ ||y||1 . Next, we
∞
show that ||fy || = |yj | = ||y||1 and that the map F : l1 → (c00 )∗ given
j=1
by F (y) = fy is a linear isometry from l1 into (c00 )∗ .
To prove F is surjective, we consider f in (c00 )∗ and let y =
(f (e1 ), f (e2 ), . . .). Next we define for n = 1, 2, . . .
,
sgn yj if 1 ≤ j ≤ n
xnj =
0 if j > n
n
n
so that ||f || ≥ f (xn ) = xnj yj = |yj |, n = 1, 2, . . .
j=1 j=1
n
so that y ∈ l1 . If x ∈ c00 then x = xi e i
i=1
for some n and hence
n
n
f (x) = xj f (ej ) = xi yi = fy (x).
j=1 i=1
6.1.5 Theorem
A
Ex Ey
∏Ex(x) ∏Ey(y)
A**
E x** E y**
Fig. 6.1
226 A First Course in Functional Analysis
6.1.6 Example
Let Ex = c00 = Ey , with the norm || · ||∞ . Then, by 6.1.4, Ex∗ is linearly
∗∗
isometric to l1 and by 6.1.3. EG
x is linearly isometric to l∞ . The completion
of c00 (that is the closure of c00 in (c∗∗00 ) is linearly isometric to c0 . Let
A ∈ (c00 → c00 ). Then A∗ can be thought of as a norm preserving linear
extension of A to l∞ .
We next explore the difference between the null spaces and the range
spaces of A, A∗ respectively.
6.1.7 Theorem
Let Ex and Ey be normed linear spaces and A ∈ (Ex → Ey ). Then
In the above, N (A) denotes the null space of A. R(A) denotes the range
space of A, N (A∗ ) and R(A∗ ) will have similar meanings.
Proof: (i) Let x ∈ Ex . Let f ∈ Ex∗ and φ ∈ Ey∗ .
Then A∗ φ(x) = f (x) = φ(Ax).
Therefore, Ax = 0 if and only f (x) = 0 ∀ f ∈ R(A∗ ).
(ii) Let φ ∈ Ex∗ Then A∗ φ = 0 if and only if φ(Ax) = A∗ φ(x) = 0 for
every x ∈ Ex .
Now, A∗ is one-to-one, that is, N (A∗ ) = {θ} if and only if φ = θ
wherever φ(y) = 0 for every y ∈ R(A). Hence, by theorem 5.1.5, this
happens if and only if the closure of R(A) = Ey , i.e., R(A), is dense in Ey .
(iii) Let y ∈ R(A) and y = Ax for some x ∈ Ex . If φ ∈ N (A∗ ) then
φ(y) = φ(Ax) = f (x) = A∗ φ(x) = 0.
Hence R(A) ⊂ {y ∈ Ey : φ(y) = 0 for all φ ∈ N (A∗ )}.
If equality holds in this inclusion, then R(A) is closed in Ey . Since
R(A) = ∩{N (φ) : φ ∈ N (A∗ )}, and each N (φ) is a closed subspace of Ey .
Conversely, let us assume that R(A) is closed in Ey . Let y0 ∈ R(A), then
by 5.1.5 there is some φ ∈ Ey∗ such that φ(y0 ) = 0 but φ(y) = 0 for every
y ∈ R(A). In particular, A∗ (φ)(x) = f (x) = φ(Ax) = 0 for all x ∈ Ex i.e.,
φ ∈ N (A∗ ). This shows that y0 ∈ {y ∈ Ey : φ(y) = 0 for all φ ∈ N (A∗ )}.
Thus, equality holds in the inclusion mentioned above.
(d) Let f ∈ R(A∗ ) and f = A∗ φ for some φ ∈ Ey∗ . If x ∈ N (A), then
f (x) = A∗ φ(x) = φ(Ax) = φ(0) = 0. Hence, R(A∗ ) ⊂ {f ∈ Ex : f (x) = 0,
for all x ∈ N (A)}.
Space of Bounded Linear Functionals 227
Note 6.2.1. Let BV ([a, b)] denote the linear space of ( )-valued 4+
functions of bounded variation on [a, b]. For w ∈ BV ([a, b]) consider
Thus || · || is a norm on BV ([a, b]). For a fixed w ∈ BV ([a, b]), let us define
4+
fw : C[a, b] → ( ) by
b
fw (x) = xdw. x ∈ C([a, b]).
a
Then fw ∈ C ∗ ([a, b]) and ||fw || ≤ ||w||. However, ||fw || may not be equal
to ||w||. For example, if z = w + 1, then fz = fw , but ||z|| = ||w|| + 1, so
that either ||fw || =
||w|| or ||fz || = ||z||.
This shows that distinct functions of bounded variation can give rise to
the same linear functional on C([a, b]). In order to overcome this difficulty
a new concept is introduced.
Space of Bounded Linear Functionals 229
a = t0 t1 s1 t2 s2 t3 tn = b
Fig. 6.2
+ |w(sj−1 ) − w(t+
j−1 )|, j = 2, . . . (n − 2).
Since the above is true for every partition Pn of [a, b], Var(y) ≤
Var(w) +
. As
> 0 is arbitrary, Var(y) ≤ Var(w). In particular, y is
of bounded variation on [a, b]. Hence y ∈ N BV [a, b].
Next, let x ∈ C([a, b]). Apart from the subtraction of the constant w(a),
the function y agrees with the function w, except possibly at the points of
discontinuities of w. Since these points are countable, they can be avoided
while calculating the Riemann-Stieljes sum
n
x(tj )[w(tj ) − w(tj−1 )],
j=1
b
n
which approximates xdw, since each sum is equal to x(tj )[y(tj ) −
a j=1
b b b
y(tj−1 )] and is approximately equal to xdy. Hence xdw = xdy.
a a a
To prove the uniqueness of y, let y0 ∈ N BV ([a, z]) be such that
b b
xdw = xdy for all x ∈ C([a, b]) and z = y − y0 . Thus z(a) =
a a
y(a) − y0 (a) = 0 − 0 = 0.
b b b
Also, since z(b) = z(b) − z(a) = dz = dy − dy0 = 0.
a a a
Now, let ξ ∈]a, b[. For a sufficiently small positive h, let
⎧
⎪
⎨ 1 if a ≤ t ≤ ξ
t−ξ
x(t) = 1− if ξ < t ≤ ξ + h
⎪
⎩ h
0 if ξ + h < t ≤ b.
Then x ∈ C([a, b]) and (x(t)) ≤ 1 for all t ∈ [a, b].
b b b
Since 0= xdy − xdy0 = xdz
a a a
ξ
t−ξ ξ+h
= dz + 1− dz
a ξ h
ξ ξ+h
t−ξ
we have z(ξ) = dz = − 1− dz.
a ξ h
It follows that |z(ξ)| ≤ Varξ ξ + h,
6.2.4 Theorem
Let E = C([a, b]). Then E is isometrically isomorphic to the subspace
of BV ([a, b]), consisting of all normalized functions of bounded variation.
If y is such a normalized function (y ∈ N BV ([a, b])), the corresponding f
is given by
b
f (x) = x(t)dy(t). (6.3)
a
Therefore, ||g|| = Var(y). Since by lemma 6.2.3 there is just one normalized
function, corresponding to the functional g a one-to-one correspondence
exists between the set of all linear functionals of C ∗ ([a, b]) and the set of
all elements of N BV ([a, b]).
It is evident that the sum of functions y1 , y2 ∈ N BV ([a, b]) corresponds
to the sum of functionals g1 , g2 ∈ C ∗ ([a, b]) and the function λy corresponds
to the functional λg, in case the functionals g1 , g2 correspond to the
functions y1 , y2 ∈ N BV ([a, b]). It therefore follows that the association
between C ∗ ([a, b]) and the space of normalized functions of bounded
variation (N BV ([a, b]) is an isomorphism. Furthermore, since
4+
where α : [0, ∞] → ( ) is of bounded variation on every subinterval of
[0, ∞). If we put t = e−u and put s = n, a positive integer, the above
integral gives rise to the form
1
μ(n) = tn dz(t), n = 0, 1, 2, (6.5)
0
the frequency density function is given by the following step function [fig.
6.3]
1
—–—
b −a
a b
Fig. 6.3
|μ(n)| ≤ Var(z), n = 0, 1, 2, . . .
In order to prove the above we need to show that for every continuous
function f ∈ C([0, 1]) and
> 0 there is a polynomial p ∈ P([0, 1]) such
that
max {|f (x) − p(x)| <
}.
x∈[0,1]
n n!
In what follows, we denote by , where n is a positive integer
k k!(n − k)!
and k an integer such that 0 < k ≤ n. The polynomial Bn (f )(x) defined
by
n
n k k
Bn (f )(x) = x (1 − x) n−k
f (6.8)
k n
k=0
is called the Bernstein polynomial associated with f . We prove our theorem
by finding a Bernstein polynomial with the required property. Before we
take up the proof, we mention some identities which will be used:
n
n k
(i) x (1 − x)n−k = [x + (1 − x)]n = 1 (6.9)
k
k=0
n
n k
(ii) x (1 − x)n−k (k − nx) = 0 (6.10)
k
k=0
(6.10) is obtained by differentiating both sides of (6.9) w.r.t. x and
multiplying both sides by x(1 − x).
On differentiating (6.10) w.r.t. x, we get
n
n
[−nxk (1 − x)n−k + xk−1 (1 − x)n−k−1 (k − nx)2 ] = 0
k
k=0
Using (6.9), (6.10) reduces to
n
n k−1
x (1 − x)n−k−1 (k − nx)2 = n (6.11)
k
k=0
Multiplying both sides by x(1 − x) and dividing by n2 , we obtain,
n 2
n k k x(1 − x)
(iii) x (1 − x) n−k
−x = (6.12)
k n n
k=0
(6.12) is the third identity to be used in proving the theorem.
It then follows from (6.8) and (6.9) that
n
n k k
f (x) − Bn (f )(x) = x (1 − x)n−k f (x) − f (6.13)
k n
k=0
n
n
k
or |f (x) − Bn (f )(x)| ≤ x (1 − x)
k n−k
f (x) − f (6.14)
k n
k=0
Since f is uniformly continuous on [0,1], we can find a
Space of Bounded Linear Functionals 235
⎫
k
δ > 0 and M s.t. x − < δ ⇒ f (x) − f k <
⎪
⎬
n n 2
⎪ (6.15)
⎭
and |f (x) < M for x ∈ [0, 1].
Let us partition
the sum on the RHS of (6.13) into two parts, denoted
stands for the sum for which x − nk < δ (x is fixed
by and .
but arbitrary, and is the sum of the remaining terms.
n
k
Thus, = xk (1 − x)n−k f (x) − f
k n
k
|x− nx |<δ
n
n k
< x (1 − x)n−k = . (6.16)
2 k 2
k=0
We next show that if n is sufficiently large then can be made less
than 2 independently of x. Since f is bounded using (6.15),
n
we get ≤ 2M xk (1 − x)n−k where the sum is taken for all
k
k
k s.t., x − ≥ δ
n
x(1 − x)
(6.12) yields δ 2 ≤
n
1 1
or ≤ 2 since max x(1 − x) = for x ∈ [0, 1]
4δ n 4
n
where = xk (1 − x)n−k ,
k
k,|x− n |>δ
k
M 2M δ 2
taking n> , < = .
δ2
4M δ 2 2
n
n k k
n−k
Hence, |f (x) − Bn (f )(x)| ≤ x (1 − x) f (x) − f
k n
k=0
< + =
.
2 2
n
k n k
Bn (f )(x) = f x (1 − x)n−k , x ∈ [0, 1], n = 0, 1, 2
n k
k=0
We express Bn (f ) as a linear combination of p0 , p1 , p2 , . . .
n−k j
n−k n−k
n−k
Since (1 − x)n−k = (−1)j x = (−1)j pj (x),
j=0
j j=0
j
6.2.11 Lemma
Let h be a linear functional on P([0, 1]) with μ(n) = h(pn ) for n =
0, 1, 2, . . . . If fkl (x) = xk (1 − x)l for x ∈ [0, 1], then
4+
(iii) The linear functional h : P([0, 1]) → ( ) defined by h(λ0 p0 +
λ1 p1 + · · · + λn pn ) = λ0 μ(0) + · · · + λn μ(n), is continuous, where n = 0, 1, 2
4+
and λ0 , λ1 , . . . , λn ∈ ( ).
Further, there is a non-decreasing function on [0,1] whose nth moment
is μ(n) if and only if dn,k ≥ 0 for all n = 0, 1, 2, . . . and k = 0, 1, 2, . . . , n.
This can only happen if and only if the linear functional h is positive.
Proof: (i) ⇒ (ii). Let z ∈ BV ([0.1]) be such that the nth moment of z is
μ(n), n = 0, 1, 2, . . . . Then
1
h(p) = pdz, z ∈ P([0, 1]),
0
define a linear functional h on P([0, 1]) such that h(pn ) = μ(n), for
n = 0, 1, 2, . . .. By lemma 6.2.11,
n
dn,k = (−1)n−k Δn−k (μ)(k)
k
n
= h(fk,n−k ) for n = 0, 1, 2, . . .
k
238 A First Course in Functional Analysis
(iii) ⇒ (ii) Since P([0, 1]) is dense in C([0, 1]) with the sup norm || ||∞ ,
there is some F ∈ C ∗ ([0, 1]) with F |P([0,1]) = h and ||F || = ||h||.
By Riesz representations theorem for C([0, 1]) there is some z ∈
N BV ([0, 1]) such that
1
F (f ) = f dz, f ∈ C([0, 1]).
0
In particular, for n = 0, 1, 2, . . .
1
μ(n) = h(pn ) = F (pn ) = tn dz(t),
0
3. Show that for any g ∈ BV ([a, b]) there is a unique g ∈ BV ([a, b]),
continuous from the right, such that
b b
f dg = f dg for all f ∈ C([a, b]) and Var (g) ≤ Var (g).
a a
1
5. Let y ∈ N BV ([a, b]) and μ(n) = 0
tn dy(t), for n = 0, 1, 2, . . .
Then show that
n
n
Var (y) = sup{ |Δn−k (μ)(k)|; n = 0, 1, 2, . . .}
k
k=0
6.3.2 Theorem
If a sequence {fn } of functionals weakly converges to itself, then {fn }
converges weakly to some linear functional f0 .
For notion of pointwise convergence see 4.5.2. Theorem 4.4.2 asserts
that E ∗ is complete, where E ∗ is the space conjugate to the normed linear
space E. Therefore if {fn } ∈ E ∗ is Cauchy, {fn } → f0 ∈ E ∗ . Therefore,
for every x ∈ E fn (x) → f0 (x) as n → ∞.
6.3.3 Theorem
Let {fn } be a sequence of bounded linear functionals defined on the
Banach space Ex .
A necessary and sufficient condition for {fn } to converge weakly to f
as n → ∞ is
(i) {||fn ||} is bounded
(ii) fn (x) → f (x) ∀ x ∈ M where the subspace M is everywhere dense
in Ex .
Proof: Let fn → f weakly, i.e., fn (x) → f (x) ∀ x ∈ Ex .
It follows from theorem 4.5.6 that {||fn ||} is bounded. Since M is a
subspace of Ex , condition (ii) is valid.
We next show that the conditions (i) and (ii) are sufficient. Let {||fn ||}
be bounded. Let L = sup ||fn ||.
Let x ∈ Ex . Since M is everywhere dense in Ex , ∃ x0 ∈ M s.t. given
arbitrary
> 0,
||x − x0 || <
/4M.
condition (ii) yields, for the above
> 0, ∃ n > n0 depending on
s.t.
|fn (x0 ) − f (x0 )| < .
2
The linear functionals f is defined on M . Hence by Hahn-Banach
extension theorem (5.1.3), we can extend f from M to the whole of Ex .
weakly
i.e., fn f.
n→∞
6.3.4 Application to the theory of quadrature formula
Let x(t) ∈ C([a, b]). Then
b
f (x) = x(t)dt (6.17)
a
The relation f (x) fn (x) which becomes an equality for all polynomials
of degree less then equal to n, is called a quadrature formula. For example,
(n)
in the case of Gaussian quadrature: t1 = a and the last element = b.
(n)
tk , (k = 1, 2, . . . n) are the n roots of Pn (t) = 2n1n! (t2 − 1)n = 0.
Consider the sequence of quadrature formula:
f (x) fn (x), n = 1, 2, 3, . . .
The problem that arises is whether the sequence {fn (x)} converges to the
value of f (x) as n → ∞ for any x(t) ∈ C([0, 1]). The theorem below answers
this question.
6.3.6 Theorem
The necessary and sufficient condition for the convergence of a sequence
of quadrature formula, i.e., in order that
kn 1
(n) (n)
lim Ck x(tk ) = x(t)dσ(t)
n→∞ 0
k=1
kn
(n)
holds for every continuous function x(t), is that |Ck | ≤ K = const,
k=1
must be true for every n,
Space of Bounded Linear Functionals 243
km
(m) (m)
fm (x) = Ck x(tk ). (6.21)
k=1
km
(n) (m)
m
(n)
Consequently, |fm (x)| ≤ |Ck x(tk )| ≤ |Ck | ||x||.
k=1 k=1
For later use we show that fm has the norm,
km
(m)
||fm || = |Ck |, (6.22)
k=1
i.e., ||fm || cannot exceed the right-hand side of (6.22) and equality holds if
we take an x0 ∈ C([0, 1]) s.t. |x0 (t)| < 1 on J and
3 (n)
(n) (n) 1 if Ck ≥ 0
x0 (tk ) = sgnCk = (n)
.
−1 if Ck < 0
kn
(n) (n)
kn
(n)
fn (x0 ) = Ck sgnCk = |Ck |.
k=0 k=0
kn
kn 1
(n) (n)
Hence, ||fn (x0 )|| = |Ck | = Ck = dσ = σ(1) − σ(0).
k=1 k=1 0
1
||xn − xn ||2 = (sin nπt − sin mπt)2 dt
0
1 1
= sin2 nπtdt − 2 sin nπt sin mπtdt
0 0
1
+ sin2 mπtdt = 1 if n = m.
0
Thus, {xn } does not converge strongly.
6.3.13 Theorem
In a finite dimensional space, notions of weak and strong convergence
are equivalent.
Proof: Let E be a finite dimensional space and {xn } a given sequence
w
such that xn −→ x. Since E is a finite dimensional there is a finite system
of linearly independent elements e1 , e2 , . . . , em s.t. every x ∈ E can be
represents as,
x = ξ1 e1 + ξ2 e2 + · · · + ξm em .
(n) (n) (n)
Let, xn = ξ1 e1 + ξ2 e2 + · · · + ξm em .
(0) (0) (0)
x0 = ξ1 e1 + ξ2 e2 + · · · + ξm em .
Now, consider the functionals fi such that
fi (ei ) = 1, fi (ej ) = 0 i = j.
(n) (0)
Then fi (xn ) = ξi , fi (x0 ) = ξi .
But since f (xn ) → f (x0 ), for every functional f ,
(0)
then also fi (xn ) → fi (x0 ), i.e., ξi (n) → ξi .
0 0
0m 0
0 (p) 0 m
0
||xp − xq || = 0 (ξi − ξi )ei 0
(q)
≤
(p) (q)
|ξi − ξi | ||ei ||
0
0 i=j 0 i=1
→ 0 as p, q → ∞, showing that in a finite dimensional normed linear
space, weak convergence of {xp } ⇐⇒ strong convergence or {xp }.
6.3.14 Remark
There also exist infinite dimensional spaces in which strong and weak
convergence of elements are equivalent.
Let E = l1 of sequences {ξ1 , ξ2 , . . . ξn . . .}
∞
s.t. the series |ξi | converges.
i=1
We note that in l1 , strong convergence of elements implies co-
ordinatewise convergence.
246 A First Course in Functional Analysis
6.3.15 Theorem
(n) (n)
Let xn = ξ1 e1 + · · · + ξi ei + · · ·
(0) (0)
x0 = ξ1 e1 + · · · + ξi ei + · · ·
w
where xn −→ x0 as n → ∞.
Let us consider the functionals fi ∈ E ∗ such that
fi (ei ) = 1, fi (ej ) = 0 i = j.
(n) (0)
Now fi (xn ) = ξi and fi (x0 ) = ξi
Since fi (xn ) → fi (x0 ) as n → ∞,
(n) (0)
ξi → ξi , i = 1, 2, 3, . . .
If ||xn || → ||x0 || as n → ∞
0 0 0 0
0 ∞ 0 0∞ 0
0 (n) 0 0 (0) 0
0 ξi ei 0 → 0 ξi ei 0
0 0 0 0
i=1 i=1
as n → ∞.
Let us introduce in E a new norm || · ||1 as follows
0∞ 0
0 0 ||xn − xo ||
0 (n) (0) 0
||xn − x0 ||1 = 0 (ξi − ξi )ei 0 = (6.23)
0 0 1 + ||xn − x0 ||
i=0 1
Since ||xn − x0 || ≥ 0
||xn − x0 ||1 ≤ ||xn − x0 ||.
Again, since {||xn ||1 } is convergent and hence bounded, ||xn ||1 ≤ M
(say).
||xn − x0 || ≤ (1 − M )−1 ||xn − x0 ||1
Thus, ||xn − x0 ||1 ≤ ||xn − x0 ||(1 − M )−1 ||xn − x0 ||1
Thus || · ||1 and || · || are equivalent norms.
M
(6.23) yields that ||xn || ≤ = L (say).
1−M
0 0
0 ∞ 0
0 (n) 0
Hence 0 ξi ei 0 ≤ L (say).
0 0
i=1
Space of Bounded Linear Functionals 247
m ∞
(n) (n)
Let Sm = ξi ei and S= ξi ei
i=1 i=1
Let
> 0 and Sm − S <
for m ≥ m0 (
).
∞
(n) (0)
Now, xn − x0 = (ξi − ξi )ei
i=1
Let
= and for
> 0, n ≥ n0 (
), m ≥ m0 (
)
2L
(m) (0) (n) (0)
then (ξi − ξi )ei ≤ |ξi − ξi | ||ei || <
2L = · 2L =
m 2L
n
for n ≥ n0 (
), m ≥ m0 (
).
Thus, ||xn − x0 || → 0 as n → ∞, proving strong convergence of {xn }.
6.3.16 Theorem
If the sequence {xn } of a normed linear space E 3converges weakly
to
kn
(n)
x0 , then there is a sequence of linear combinations Ck xk which
k=1
converges strongly to x0 .
In other words, x0 belongs to a closed linear subspace L, spanned by
the elements x1 , x2 , . . . xn , . . . .
Proof: Let us assume that the theorem is not true, i.e., x0 does not
belong to the closed subspace L. Then, by theorem 5.1.5, there is a linear
functional f ∈ E ∗ , such that f (x0 ) = 1 and f (xn ) = 0, n = 1, 2, . . ..
But this means that f (xn ) does not converge to f (x0 ), contradicting the
w
hypothesis that xn −→ x0 .
6.3.17 Theorem
Let A be a bounded linear operator with domain Ex and range in Ey ,
both normed linear spaces. If the sequence {xn } ⊂ Ex converges weakly to
x0 ⊂ Ex , then the sequence {Axn } ⊂ Ey converges weakly to Ax0 ∈ Ey .
Proof: Let φ ∈ Ey∗ be any functional. Then φ(Axn ) = f (xn ), f ∈ Ex∗ .
Analogously φ(Ax0 ) = f (x0 ).
w
Since xn −→ x0 , f (xn ) → f (x0 ) i.e., φ(Axn ) → φ(Ax0 ). Since φ is
w
an arbitrary functional in Ey∗ , it follows that Axn −→ Ax0 . Thus, every
bounded linear operator is not only strongly, but also weakly continuous.
6.3.18 Theorem
If a sequence {xn } in a normed linear space converges weakly to x0 ,
then the norm of the elements of this sequence is bounded.
We regard xn (n = 1, 2, . . .) as the elements of E ∗∗ , conjugate to E ∗ then
the weak convergence of {xn } to x0 means that the sequence of functions
248 A First Course in Functional Analysis
6.3.20 Theorem
In order that a sequence {xn } of a normed linear space E converges
weakly to x0 , it is necessary and sufficient that
(i) the sequence {||xn ||} is bounded and
(ii) f (xn ) → f (x0 ) for every f of a certain set Ω of linear functionals,
linear combination of whose elements are everywhere dense in E ∗ .
Proof: This theorem is a particular case of theorem 6.3.3. This is because
convergence of {xn } ⊂ E to x0 ∈ E is equivalent to the convergence of the
linear functionals {xn } ⊂ E ∗∗ to xo ∈ E ∗∗ .
Space of Bounded Linear Functionals 249
Problems
1. Let E be a normed linear space.
(a) If X is a closed convex subset of E, {xn } is a sequence in X and
w
xn −→ x in E, then prove that x ∈ X (6.3.20).
w
(b) Let Y be a closed subspace of E. If xn −→ x in E, then show
w
that xn + Y −→ x + Y in E/Y .
w
2. In a Hilbert space H, if {xn } −→ x and ||xn || → ||x|| as n → ∞,
show that {xn } converges to x strongly.
pn
3. Let fm (x) = Cnm x(tn,m ) be a sequence of quadrature formulae
m=1
b
for, f (x) = k(x, t)dt on the Banach space Ex = C([a, b]).
a
{xn } converges weakly to x. Show that the space C([a, b]) is not
weakly sequentially complete.
w
8. If xn −→ x0 in a normed linear space E, show that x0 ∈ Y , where
Y = span {xn }. (Use theorem 5.1.5).
w
9. Let {xn } be a sequence in a normed linear space E such that xn −→ x
in E. Prove that there is a sequence {yn } or linear combination of
elements of {xn } which converges strongly to x. (Use Hahn-Banach
theorem).
10. In the space l2 , we consider a sequence {Tn }, where Tn : l2 → l2 is
defined by
Tn x = (0, 0, . . . , 0, ξ1 , ξ2 , . . .), x = {xn } ∈ l2 .
Show that
(i) Tn is linear and bounded
(ii) {Tn } is weakly operator convergent to 0, but not strongly.
(Note that l2 is a Hilbert space).
11. Let E be a separable Banach space and M ⊂ E ∗ a bounded set. Show
that every sequence of elements of M contains a subsequence which
is weak∗ convergent to an element of E ∗ .
12. Let E = C([a, b]) with the sup norm. Fix t0
(a, b). For each positive
integer n with t0 + n4 < b, let
⎧ 4
⎪
⎪ 0 if a ≤ t ≤ t0 , t0 + ≤ t ≤ b
⎪
⎪ n
⎪
⎨ 2
xn (t) = n(t − t0 ) if t0 ≤ t ≤ t0 + ,
⎪
⎪ n
⎪
⎪
⎪
⎩ n 4
− t + t0
2
if t0 + ≤ t ≤ t0 + ,
4
n n n
w
Then show that xn −→ θ in E but xn (t) →
/ θ in E.
13. Let Ex be a Banach space and Ey be a normed space. Let {Fn } be
a sequence in (Ex → Ey ) such that for each fixed x ∈ Ex , {Fn (x)} is
w
weakly convergent in Ey . If Fn (x) −→ y in Ey , let F (x) = y. Then
show that F ∈ (Ex → Ey ) and
||F || ≤ lim inf ||Fn || ≤ sup ||Fn || < ∞, n = 1, 2, . . .
n→∞
6.4 Reflexivity
In 5.6.6 the notion of canonical or natural embedding of a normed
linear space E into its second conjugate E ∗∗ was introduced. We have
discussed when two normed linear spaces are said to be reflexive and some
relevant theorems. Since the conjugate spaces E, E ∗ , E ∗∗ often appear in
discussions on reflexivity of spaces, there may be some relationships between
weak convergence and reflexivity. In what follows some results which were
not discussed in 5.6 are discussed.
6.4.1 Theorem
Let E be a normed linear space and {f1 , . . . fn } be a linearly independent
subset of E ∗ . Then there are e1 , e2 , . . . , en in E such that fj (ei ) = δij for
i, j = 1, 2, . . . n.
Proof: We prove by induction on m. If m = 1, then since {f1 } is
linearly independent, let a0 ∈ E with f1 (a0 ) = 0. Let e1 = f1a(a0 0 ) .
Hence, f1 (e1 ) = 1. Next let us assume that the result is true for
m = k. Let {f1 , f2 , . . . , fk+1 } be a linearly independent subset of E ∗ .
Since {f1 , f2 , . . . fk } is linearly independent, there are a1 , a2 , . . . ak in E
such that fj (ai ) = δij for 1 ≤ i, j ≤ k. We claim that there is some a0 ∈ E
such that fj (a0 ) = 0 for 1 ≤ j ≤ k but fk+1 (a0 ) = 0. For x ∈ E, let
a0 = f1 (x)a1 + f2 (x)a2 + · · · + fk (x)ak .
-
k -
k
Then (x − a) ∈ N (fj ). If N (fj ) ⊂ N (fk+1 ),
j=1 j=1
a0 fj (a0 )
fj (ek+1 ) = fj = =0
fk+1 (a0 ) fk+1 (a0 )
for j = 1, 2, . . . , k, since fj (a0 ) = 0.
Also fk+1 (ek+1 ) = 1.
Hence fj (ei ) = δij , i, j = 1, 2, . . . k + 1
for all h1 , h2 , . . . , hm in 4 +
( ).
(b) Let S be a finite dimensional subspace of E ∗ and Fx ∈ E ∗∗ . If
> 0,
then there is some x ∈ E such that
F S = φ(x )S and ||x || < ||Fx || +
.
Proof: (a) Suppose that for every
> 0, there is some x ∈ E such that
fj (x ) = kj for each j = 1, . . . , m and ||x || < α+
. Let us fix h1 , h2 , . . . , hm
in 4 + ( ). Then
⎛ ⎞
m m m
h k = h f (x )
= ⎝ h f ⎠ (x )
j j j j j j
j=1 j=1 j=1
0 0 0 0
0m 0 0 0
0 0 0m 0
≤00 h f
j j0
0 ||x || < (α +
) 0
0 h f 0
j j 0.
0j=1 0 0 j=1 0
As this is true for every
> 0, we conclude that
0 0
0 0
m 0m 0
h k ≤ α 0 h f 0
j j 0.
j j 0
j=1 0 j=1 0
m
Conversely, suppose that for all h1 , h2 , . . . , hm in 4 +
( ), hj kj ≤
j=1
0 0
0 0
0m 0
α0 h 0
j j 0. It may be noted thats {f1 , f2 , . . . , fm } can be assumed to be
f
0
0j=1 0
a linearly independent set. If that is not so, let f1 , f2 , . . . , fn with n ≤ m
254 A First Course in Functional Analysis
4 +
Consider the map F : E → m ( m ) given by F(x) = (f1 (x), . . . , fm (x)).
Clearly, F is a linear map. Next, we show that it is a surjective (onto)
4
mapping. To this end consider (h1 , h2 , . . . , hm ) ∈ m (or m ). Since +
{f1 , f2 , . . . , fm } are linearly independent, it follows from theorem 6.4.1 that
there exist e1 , e2 , . . . , em in E such that fj (ei ) = δij , 1 ≤ i, j ≤ m. If we
take x = h1 e1 + · · · + hm em , then it follows that F(x) = (h1 , h2 , . . . , hm ).
We next want to show that F maps each open subset of E onto an open
4 +
subset of m (or m ). Since F is non-zero we can find a non-zero vector,
4
‘a’ in E s.t. F(a) = (1, 1, . . . 1) ∈ m (Cm ). Let P be an open set in E.
Then there exists an open ball U (x, r) ⊂ E with x ∈ E and r ∈ . 4
We can now find a scalar k such that
r
x − ka ∈ U (x, r) where 0 < |k| < .
||a||
Hence, x − ka ∈ P with the above choice of k.
Therefore, F(x − ka) = F(x) − kF(a) = F(x) − k ∈ F(E). Thus
,
4+
k ∈ ( ) : F(x) = |k | <
r
||a||
⊂ F (E),
6.4.3 Remark
(i) It may be noted if we restrict ourselves to a finite dimensional
subspace of E, then we are close to reflexivity.
The relationship between reflexivity and weak convergence is
demonstrated in the following theorem.
6.4.4 Theorem (Eberlein, 1947)
Let E be a normed linear space. Then E is reflexive if and only if every
bounded sequence has a weakly convergent subsequence.
Proof: For proof see Limaye [33].
6.4.5 Uniform convexity
We next explore some geometric condition which implies reflexivity. In
2.1.12 we have seen that a closed unit ball of a normed linear space E is a
convex set of E. In the case of the strict convexity of E, the mid-point of
the segment joining two points on the unit sphere of E does not lie on the
256 A First Course in Functional Analysis
6.4.7 Lemma
A uniformly convex space is strictly convex. This is evident from
the definition itself.
6.4.8 Lemma
If E is finite dimensional and strictly convex, then E is uniformly
convex.
Proof: For
> 0, let
From theorem 4.4.2, we can conclude that the space of bounded linear
functionals defined on a normed linear space E is complete and hence a
Banach space. Thus the dual E ∗ and in turn the second dual E ∗∗ of
the normed linear space E are Banach spaces. Since E is reflexive, E
is isometrically isomorphic to E ∗∗ and hence E is a Banach space. Also, in
any equivalent norm on E, the dual E ∗ and the second dual E ∗∗ remains
unchanged, so that E remains reflexive.
Hence, we can assume without loss of generality that E is a uniformly
convex Banach space in the given norm || || on E.
Let Fx ∈ E ∗∗ . Without loss of generality we assume that ||Fx || = 1.
To show that there is some x ∈ E with φ(x) = Fx φ : E → E ∗∗ being
a canonical embedding. First, we find a sequence {fn } in E ∗ such that
||fn || = 1 and |Fx (fn )| > 1 − n1 for n = 1, 2, . . ..
6.4.13 Remark
The converse of Milman’s theorem is false [see Limaye[33]].
Problems
1. Let E be a reflexive normed linear space. Then show that E is strictly
convex (resp. smooth) if and only if E ∗ is smooth (resp. strictly
convex).
(Hint: A normed linear space E is said to be smooth if, for every
x0 ∈ E with ||x0 || = 1, there is a unique supporting hyperplane [see
4.3.7] for B(θ, 1) at x0 .)
2. [Weak Schauder basis. Let E be a normed linear space. A
countable subset {a1 , a2 , . . .} of E is called a weak Schauder basis
for E if ||ai || = 1 for each i and for every x ∈ E, there are unique
n
4+
αi ∈ ( ) i = 1, 2, . . . such that
w
αi ai −→ x as n → ∞.
i=1
Weak∗ Schauder bases. A countable subset {f1 , f2 , . . .} of E ∗
is called a Weak∗ Schauder basis if ||fi || = 1 for all i and for
4+
every g ∈ E ∗ , there are unique βi ∈ ( ), i = 1, 2, . . . such that
n
w
βi fi −→ g as n → ∞.]
i=1
Let E be a reflexive normed linear space and {a1 , a2 , . . .} be a
Schauder basis for E with coefficient functionals {g1 , g2 , . . .}. If
fn = gn /||gn ||, n = 1, 2, . . . then show that {f1 , f2 , . . .} is a Schauder
basis for E ∗ with coefficient functionals (||g1 ||Fa1 , ||g2 ||Fa2 , . . .).
3. Let E be a separable normed linear space. Let {xn } be a dense subset
of {x ∈ E : ||x|| = 1}.
(c) Show that l1 is strictly convex but not reflexive in some norm
which is equivalent to the norm || ||1 .
4. Let E be a uniformly convex normed linear space, x ∈ E, and {xn }
be a sequence in E.
(a) If ||x|| = 1, ||xn || → 1 and ||xn + x|| → 2 then show that
||xn − x|| → 0.
w
(b) Show that xn → x in E if and only if xn −→ x in E and
lim sup ||xn || ≤ ||x||.
n→∞
ψ(α1 , α2 , . . . , αn ) = ||α1 x1 + α2 x2 + · · · + αn xn ||
Therfore m > 0.
Given an arbitrary K > 0,
φ(α1 , α2 , . . . αn ) ≥ ||α1 x1 + α2 x2 + · · · + αn xn || − ||x||
262 A First Course in Functional Analysis
0 0
n 0α x + · · · + α x 0
0 0
= 2 1 1 n n
αi · 0 +n 0 − ||x||
0 i=1 iα 2 0
i=1
n
≥ αi2 · m − ||x||,
i=1
,
α1 α2 αn
since 2 , 2 , · · · , 2 lie on a unit ball.
i αi i αi i αi
n
2
Thus if α > 1 (K + ||x||), φ(α1 , α2 , . . . αn ) > K
i
i=1
m
6.5.2 Theorem
(0) (0) (0)
There exist real numbers α1 , α2 , . . ., αn , such that φ(α1 , α2 , . . .,
(0)
αn ) = ||x−α1 x1 −α2 x2 . . . αn xn || assumes its minimum for α1 = α1 , α2 =
(0) (0)
α2 . . . αn = αn .
Proof: If x depends linearly on x1 , x2 , . . . , xn , then the theorem is true
immediately. Let us assume that x does not lie in the subspace spanned by
x 1 , x2 , . . . , xn .
We first show that φ(α1 , α2 , . . . , αn ) is a continuous function of its
arguments.
6.5.3 Remark
n
(0)
(i) The linear combination λi xi , giving the best approximation of
i=1
the element x, is in general not unique.
(ii) Let Y be a finite dimensional subspace of C([0, 1]). Then the best
approximation out of Y is unique for every x ∈ C([a, b]) if and only if Y
satisfies the Haar condition [see Kreyszig [30]].
(iii) However, there exist certain spaces in which the best approximation
is everywhere uniquely defined.
6.5.4 Definition: strictly normed
A space E is said to be strictly normed if the equality ||x + y|| =
||x|| + ||y|| for x = θ, y = θ, is possible only when y = ax, with a > 0.
6.5.5 Theorem
In a strictly normed linear space the best approximation of an arbitrary
x in terms of a linear combination of a given finite system of linearly
independent elements is unique.
n
Proof: Let us suppose that there exist two linear combinations α i xi
i=1
n
and βi xi such that
i=1
0 0 0 0
0 n 0 0 n 0
0 0 0 0
0x − α i xi 0 = 0x − βi xi 0 = d,
0 0 0 0
i=1 i=1
0 0
0 n 0
0 0
where d = min 0x − ri xi 0 > 0,
ri 0 0
i=1
0 0 0 0 0 0
0 α i + βi 0 10 0 10 0
n n n
0 0 0 0 0 0
then 0x − x i 0 ≤ 0x − α i x i 0 + 0x − β i xi 0
0 2 0 2 0 0 20 0
i=1 i=1 i=1
1 1
= d + d = d.
0 0 2 2
0 n 0
0 αi + βi 0
and since 0x − xi 0 ≥ d,
0 2 0
i=1
0 0
0 n
α i + βi 0
0 0
we have 0x − xi 0 = d,
0 2 0
i=1
0 0 0
0
0 n
αi + βi 0 0 n 0
0 0 01 0
Consequently, 0x − xi 0 = 0 x− α1 xi 0
0 2 0 02 0
i=1 i=1 0
0
01 n 0
0 0
+0 x− βi xi 0.
02 0
i=0
264 A First Course in Functional Analysis
6.5.7 Lemma
Let E be a reflexive normed linear space and M be a non-empty closed
convex subset of E. Then, for every x ∈ E, there is some y ∈ M such that
||x − y|| = dist (x, M ), that is there is a best approximation to x from M .
Proof: Let x ∈ E and d = dist (x, M ). If x ∈ M then the result is
trivially satisfied. Let x ∈ M . Then there is a sequence {yn } in M
such that ||x − yn || → d as n → ∞. Since x is known and {x − yn } is
bounded, {yn } is bounded, and since E is reflexive, {yn } contains a weakly
convergent subsequence {ynp }, (6.4.4). Now, {ynp } ⊂ M and M being
closed, lim φ1 (ynp ) = φ(y 1 ), y 1 ∈ M and φ1 is any linear functional.
p→∞
Therefore, lim φ1 (x − ynp ) = φ1 (x − y 1 ).
p→∞
Since {ynp } is a subsequence of {yn },
CLOSED GRAPH
THEOREM AND ITS
CONSEQUENCES
For, x1 , x2 ∈ Ex , y1 , y2 ∈ Ey ,
⎫
(x1 , y1 ) + (x2 , y2 ) = (x1 + x2 , y1 + y2 ) ⎪
⎪
⎬
For, x ∈ Ex , y ∈ Ey , (7.2)
⎪
⎪
⎭
267
268 A First Course in Functional Analysis
Since G(T ) and D(T ) are complete, we can apply the bounded inverse
Closed Graph Theorem and Its Consequences 269
theorem (theorem 7.3) and see that P −1 is bounded, say ||(x, T x)|| ≤ b||x||,
for some b and all x ∈ D(T ).
7.1.4 Remark
G(T ) is closed if and only if z ∈ G(T ) implies z ∈ G(T ). Now, z ∈ G(T )
if and only if there are zn = (xn , yn ) ∈ G(T ) such that zn → z, hence
xn → x, T xn → T x.
This leads to the following theorem, where an important criterion for
an operator T to be closed is discovered.
7.1.5 Theorem (closed linear operator)
Let T : D(T ) ⊂ Ex → Ey be a linear operator, where Ex and Ey are
normed linear spaces. Then T is closed if and only if it fulfils the following
condition: If xn → x where xn ∈ D(T ) and T xn → y together imply that
x ∈ D(T ) and T x = y.
7.1.6 Remark
(i) If T is a continuous linear operator, then T is closed.
Since T is continuous, xn → x in Ex implies that T xn → T x in Ey .
(ii) A closed linear operator need not be continuous. For example, let
Ex = Ey = 4 and T x = x1 for x = θ and T θ = θ. Here, if xn → θ, then
T xn → θ, showing that T is closed. But T is not continuous.
(iii) Given T is closed, and that two sequences, {xn } and {xn }, in the
domain converge to the same limit x; if the corresponding sequences {T xn }
and {T xn } both converge, then the latter have the same limit.
T being closed, xn → x and T xn → y1 imply x ∈ D(T ) and T x = y1 .
Since {xn } → x, T being closed, xn → x and T xn → y2 imply that
x ∈ D(T ) and y2 = T x.
Thus, {T xn } and {T xn } have the same limit.
7.1.7 Example (differential operator)
We refer to example 4.2.11.
We have seen that the operator A given by Ax(t) = x (t) where
Ex = C([0, 1]) and D(A) ⊂ Ex is the subspace of functions having
continuous derivatives, is not bounded. We show now that A is a closed
operator. Let xn ∈ D(A) be such that
xn → x and Axn = xn → y.
t t t
y(τ )dτ = lim xn (τ )dτ = lim xn (τ )dτ = x(t) − x(0),
0 0 n→∞ n→∞ 0
t
i.e., x(t) = x(0) + y(τ )dτ .
0
This shows that x ∈ D(A) and x = y. The theorem 7.1.5 now implies
that A is closed.
Note 7.1.1. Here, D(A) is not closed in Ex = C([0, 1]), for otherwise A
would be bounded by the closed graph theorem.
7.1.8 Theorem
Closedness does not imply boundedness of a linear operator. Conversely,
boundedness does not imply closedness.
Proof: The first statement is shown to be true by examples 7.1.6 (ii) and
7.1.7. The second statement is demonstrated by the following example. Let
T : D(T ) → D(T ) ⊂ Ex be the identity operator on D(T ), where D(T ) is
a proper dense subspace of a normed linear space Ex . It is evident that T
is linear and bounded. However, we show that T is not closed. Let us take
x ∈ Ex − D(T ) and a sequence {xn } in D(T ) which converges to x.
7.1.9 Lemma (closed operator)
Since a broad class of operators in mathematical and theoretical physics
are differential operators and hence unbounded operators, it is important
to determine the domain and extensions of such operators. The following
lemma will be an aid in investigation in this direction.
Let T : D(T ) → Ey be a bounded linear operator with domain
D(T ) ⊆ Ex , where Ex and Ey are normed linear spaces. Then:
(a) If D(T ) is a closed subset of Ex , then T is closed.
(b) IF T is closed and Ey is complete, then D(T ) is a closed subset of
Ex .
Proof: (a) If {xn } is in D(T ) and converges, say, xn → x and is such that
{T xn } also converges, then x ∈ D(T ) = D(T ), since D(T ) is closed.
T xn → T x since T is continuous.
Hence, T is closed by theorem 7.1.5.
(b) For x ∈ D(T ) there is a sequence {xn } in D(T ) such that xn → x.
Since T is bounded,
x = y + z, (7.6)
with y ∈ M and z ∈ N .
Thus, if E = M ⊕ N then M ∩ N = {θ}.
7.1.12 Definition: projection
A linear map P from a linear space E to itself is called a projection if
P2 = P.
7.1.13 Lemma
If a normed linear space E is the direct sum of two subspaces M and
N and if P is a projection of E onto M , then
(i) P x = x if and only if x ∈ M ;
(ii) P x = θ if and only if x ∈ N .
Proof: If x ∈ E, then x = y + z where y ∈ M and z ∈ N .
Since P is a projection of E onto M
Px = y
if P x = x, y = x and z = θ. Similarly, if x ∈ M, P x = x.
If P x = θ then y = θ and hence x ∈ N . Similarly, if x ∈ N, P x = θ.
If R(P ) and N (P ) denote respectively the range space and null space
of P , then
R(P ) = N (I − P ), N (P ) = R(I − P ).
Therefore, E = R(P ) + N (I − P ) and R(P ) ∩ N (P ) = {θ} for every
projection P defined on E.
272 A First Course in Functional Analysis
(i) Given a set A ⊆ Ex we shall write [see figs 7.1(a) and 7.1(b)]
A + g = {x ∈ Ex : x = a + g, a ∈ A, g ∈ Ex } (7.8)
and similarly for subsets of Ey .
A
αA
A+g
A
a +g
Fig. 7.1(a) Illustration of formula (7.7) Fig. 7.1(b) Illustration of formula (7.8)
1
We consider the open ball B1 = B 0, ⊆ Ex .
2
Any fixed x ∈ Ex is in kB1 with real k sufficiently large (k > 2||x||).
∞
*
Hence, Ex = kB1 .
k=1
Since T is surjective and linear,
∞
∞ ∞
* * *
Ey = T (Ex ) = T kB1 = kT (B1 ) = kT (B1 ). (7.9)
k=1 k=1 k=1
274 A First Course in Functional Analysis
∞
*
Since Ey is complete and kT (B1 ) is equal to Ey , (7.9) holds. Since
k=1
Ey is complete, it is a set of the second category by Baire’s category theorem
*∞
(theorem 1.4.20). Hence Ey = kT (B1 ) cannot be expressed as the
k=1
countable union of nowhere dense sets. Hence, at least one ball kT (B1 )
must contain an open ball. This means that T (B1 ) must contain an open
ball B ∗ = B(y0 ,
) ⊂ T (B1 ). It therefore follows from (7.8) that
vn → T zn ⊆ T (B1 ) and vn → y0 .
1
Since wn , zn ∈ B1 and B1 has radius 2, we obtain ||wn − zn ||
||wn || + ||zn || < 1.
||v − y|| < .
4
Now, v ∈ T (B1 ) implies that there is a x1 ∈ B1 such that v = T x1 .
Hence ||y − T x1 || < .
4
From the above and (7.13), putting n = 2, we see that
y − T x1 ∈ V2 = T (B2 ).
We can again find a x2 ∈ B2 such that
||(y − T x1 ) − T x2 || < .
8
Hence y − T x1 − T x2 ∈ V3 ⊆ T (B3 ) and so on.
Proceeding in the above manner we get at the nth stage, an xn ∈ Bn
such that0 0
0 n 0
0 0
0y − T xk 0 < n+1 , (n = 1, 2, . . .) (7.14)
0 0 2
k=1
writing zn = x1 + x2 + · · · + xn , and since xk ∈ Bk , so that ||xk || < 1/2k
we have, for n > m,
n n
1
||zn − zm || ||xk || <
2k
k=m+1 k=m+1
1 1 1 1
= m+1 1 + + 2 + · · · + n−m−1 → 0 as m → ∞.
2 2 2 2
Hence, {zn } is a Cauchy sequence and Ex being complete {zn } → z ∈
Ex . Also z ∈ B(0, 1) since B(0, 1) has radius 1 and
∞
∞
1
||xk || < = 1. (7.15)
2k
k=1 k=1
Since T is continuous, T zn → T z and (7.14) shows that T z = y. Hence
y ∈ T (B(0, 1)).
Let Ex = c00 with || ||1 and Ey = c00 with || · ||∞ . If P (x) = x for
x ∈ Ex , then P : Ex → Ey is bijective, linear and continuous. But P −1 is
not continuous since, for xn = (1, 1, 1, . . . , 1 , 0, 0, 0 . . .) we have ||xn ||∞ = 1
A BC D
n
and ||P −1 (xn )|| = ||xn ||1 = n for all n = 1, 2, . . ..
7.3.2 Definition: stronger norm, comparable norm
Given E a normed linear space, || · || on E is said to be stronger than
the norm || · || if for every x ∈ E and every
> 0, there is some δ > 0 such
B||·|| (x, δ) ⊆ B||·|| (x,
). Here, B||·|| (x, δ) denotes an open ball in E w.r.t.
|| · ||. Similarly B||·|| (x,
) denotes an open ball in E w.r.t. || · || . In other
words, || · || is stronger than || · || if and only if every open subset of E
with respect to || · || is also an open subset with respect to || · ||.
The norms || · || and || · || are said to be comparable if one of them is
stronger than the other. For definition of two equivalent norms, || · || and
|| · || , see 2.3.5.
7.3.3 Theorem
Let || · || and || · || be norms on a linear space E. Then the norm || · || is
stronger than ||·|| if and only if there is some α > 0 such that ||x|| ≤ α||x||
for all x ∈ E.
Proof: Let || · || be stronger than || · || , then there is some r > 0 such that
0 0
0 rx 0
then 0 0 0 < 1 , i.e., ||x|| < (1 +
) ||x||.
(1 +
)||x|| 0 r
1
Since
> 0 is arbitrary, ||x|| ≤ ||x||.
r
1
or ||x|| ≤ α||x|| with α = .
r
Conversely, let ||x|| ≤ α||x|| for all x ∈ E.
Let {xn } be a sequence in E such that ||xn − x|| → 0.
Since ||xn − x|| ≤ α||xn − x|| → 0.
Hence, the || · || is stronger than the norm || · || .
7.3.4 Two-norm theorem
Let E be a Banach space in the norm || · ||. Then a norm || · || of the
linear space E is equivalent to the norm || · || if and only if, E is also a
Banach space in the norm || · || and the norm || · || is comparable to the
norm || · ||.
Proof: If the norms || · || and || · || are equivalent, then clearly they are
comparable.
Therefore, α1 ||x|| ≤ ||x|| ≤ α2 ||x||, α1 , α2 ≥ 0 for all x ∈ E. Let {xn }
be a Cauchy sequence in the Banach space E with norm || · ||.
Then, if xn → x in E with norm || · ||,
7.4.1 Theorem
The functionals fn = αn (x) for a given x ∈ E are bounded.
We consider the vector space E of all sequence (α1 ,0α2 , . . . , α0n , . . .) for
∞
0n 0
0 0
which αn en converges in E. The norm ||y|| = sup 0 αi ei 0 converts
n 0 0
n=1 i=1
E into a normed vector space. We show that E is a Banach space. Let
(m)
ym = {αn }, m = 1, 2, . . . be a Cauchy sequence in E . Let
> 0, then
there is a N , such that m, p > N implies
0 0
0n 0
0 (m) (p) 0
||ym − yp || = sup 0 (αi − αi )ei 0 <
.
n 0 0
i=1
(m) (p)
But this implies ||αn
− αn ||
< 2
for every n. Hence for every n,
lim αn(m) = αn exists. It remains to be shown that
m→∞
y = (α1 , α2 , . . . , αn , . . .) ∈ E and lim yn = y.
n→∞
(m) (m) (m)
Now, ym = {α1 , α2 , . . . αn . . .} ∈ E .
Since in E convergence implies coordinatewise convergence and since
lim αnm = αn , lim ym = (α1 , α2 , . . . αn , . . .).
n→∞ n→∞
0 0
0n 0
0 0
Now, ||y − ym || = sup 0 (αi − αi )ei 0 → 0 as m → ∞.
m
n 0 0
i=1
Hence y = (α1 , α2 , . . . αn , . . .) ∈ E and {ym } being Cauchy, E is a
Banach space.
Let us next consider a mapping P : E → E for which y = (α1 , α2 , . . .) ∈
∞
E such that Py = αn en ∈ E. If z = (β1 , β2 , . . .) ∈ E such that
n=1
Closed Graph Theorem and Its Consequences 281
∞
∞
Pz = βn en ∈ E, then P (y + z) = (αn + βn )en = P y + P z showing
n=1 n=1
P is linear. Since {ei } are linearly independent P y = θ ⇐⇒ y = θ. Hence
P is one-to-one. Now {en } being a Schauder basis in E, every element in
∞
E is representable in the form rn en where (r1 , r, . . . rn . . .) ∈ E . Hence
n=1
P is onto. P is bounded since
0 0 0 0
0 n 0 0n 0
0 0 0 0
sup 0 αi ei 0 ≥ lim 0 αi ei 0 .
n 0 0 n 0 0
i=1 i=1
COMPACT
OPERATORS ON
NORMED LINEAR
SPACES
282
Compact Operators on Normed Linear Spaces 283
Ex to a bounded subset of Ey .
(ii) A compact linear operator A is stronger than a bounded linear
operator in the sense that A(B(0, 1)) is a compact subset of Ey given B(0, 1)
an open unit ball.
(iii) A compact linear operator is also known as a completely
continuous operator in view of a result we shall prove in 8.1.14(a).
8.1.3 Remark
(i) A compact linear operator is continuous, but the converse is not
always true. For example, if Ex is an infinite dimensional normed linear
space, then the identity map I on Ex is clearly linear and continuous, but
it is not compact. See example 1.6.16.
8.1.4 Lemma
Let Ex and Ey be normed linear spaces.
Further, A1 , A2 map B(0, 1) into A1 (B(0, 1)) and A2 (B(0, 1)) which are
respectively compact.
Then (i) (A1 + A2 )(B(0, 1)) is compact.
(ii) A1 A2 (B(0, 1)) is compact.
Proof: Let ||xn || ≤ 1 and {xn } is a Cauchy sequence. Since A1 (B(0, 1))
is compact and is hence sequentially compact (1.6), {A1 xn } contains a
convergent subsequence {A1 xnp }.
A2 being compact, we can similarly argue that {A2 xnq } is convergent.
Let {xnr } be a subsequence of both {xnp } and {xnq }.
Then {xnr } is Cauchy.
Moreover, (A1 + A2 ) (xnr ) is convergent.
Hence (A1 + A2 ) is compact.
(ii) Let ||xn || ≤ 1 and {xn } is convergent in Ex . {A2 xn } being
compact and sequentially compact (1.6.17) {A2 xn } contains a convergent
subsequence {A2 xnp } ⊆ Ey . Hence {A2 xnp } is bounded. A1 being compact
(A1 A2 )(xnp ) is a compact sequence. Hence, A1 A2 maps bounded sequence
{xn } into a compact sequence. Hence A1 A2 is compact.
8.1.5 Examples
(1) Let Ex = Ey = C([0, 1]) and let
1
Ax = y(t) = K(t, s)x(s)ds.
0
8.1.8 Theorem
Let A be a linear compact operator mapping an infinite dimensional
space E into itself and let B be an arbitrary bounded linear operator acting
in the same space. Then AB and BA are compact.
Proof: See lemma 8.1.4.
Note 8.1.1. In case, A is a compact linear operator mapping a linear
space E → E and admits of an inverse A−1 , then A · A−1 = I. Since I is
not compact, A−1 is not bounded.
8.1.9 Theorem
If a sequence {An } of compact linear operators mapping a normed linear
space Ex into a Banach space Ey converges strongly to the operator A, that
is if ||An − A|| → 0, then A is also a compact operator.
Proof: Let M be a bounded set in Ex and α a constant such that ||x|| ≤ α
for every x ∈ M . For given
> 0, there is an index n0 such that
||An − A|| <
/α, for n ≥ n0 (
). Let A(M ) = L and An0 (M ) = N .
We assert that the set An0 (M ) = N is a finite
-net of L. Let us take
for every y ∈ L one of the pre-images x ∈ M and put y0 = An0 x ∈ N ,
to receive ||y − y0 || = ||Ax − An0 x|| ≤ ||A − An0 || ||x|| <
/α · α =
.
On the other hand, since An0 is compact and M is bounded, the set N
is compact. It follows then L for every
> 0 has a compact
-net and is
therefore itself compact (theorem 1.6.18). Thus, the operator A maps an
arbitrary bounded set into a set whose closure is compact set and hence
the operator A is compact.
8.1.10 Example
1
1. If Ex = Ey = L2 ([0, 1]), then the operator, Ax−y = K(t, s)x(s)ds
1 1 0
12 12
1 1 1
|y(t)| = K(t, s)x(s)ds ≤ 2
K (t, s)ds 2
x (s)ds ≤ Lα
0 0 0
8.1.11 Remark
The limit of a weakly convergent sequence {An } of compact operators is
not necessarily compact.
Let us consider an infinite dimensional Banach space E with a basis
{ei }. Then every x ∈ E can be written in the form
∞
x= ξi ei .
i=1
n
Let Sn x = ξi ei where Sn x is a projection of x to a finite dimensional
i=1
space.
Let us consider the unit ball B(0, 1) = {x : x ∈ E, ||x|| ≤ 1}.
Then Sn (B(0, 1)) is closed and bounded in the n-dimensional space En
and hence compact.
Thus, Sn is compact.
w w
As n → ∞, Sn x −→ x or Sn −→ I, where the identity operator I is
not compact.
8.1.12 Theorem (Schauder, 1930) [49]
Let Ex and Ey be normed linear spaces and A ∈ (Ex → Ey ). If A is
compact then A∗ is a compact linear operator mapping Ey∗ into Ex∗ . The
converse holds if Ey is a Banach space.
Proof: Let A be a compact linear operator mapping Ex into Ey . Let us
consider a bounded sequence {φn } in Ey∗ . For y1 , y2 ∈ Ey
For i, j = 1, 2, . . ., we have
||A∗ (φni ) − A∗ (φnj )|| = sup{|A∗ (φni − φnj )(x)| : ||x|| ≤ 1}
= sup{|(φni − φnj )(Ax)| : ||x|| ≤ 1}
≤ sup{|φni (y) − φnj (y)| : y ∈ L}.
8.1.15 Theorem
8.1.16 Remark
8.1.17 Theorem
8.2.2 Remark
The above result shows that if A is a compact linear operator mapping
Ex into Ex , and if {xn } is a bounded sequence of approximate solution of
Ax − kx = y, then a subsequence of {xn } converges to an exact solution
of the above equation. The following result, which is based on Riesz
lemma (2.3.7), is instrumental in analysing the spectrum of a compact
operator.
8.2.3 Lemma
Let Ex be a normed linear space and A : Ex → Ex .
4
(a) Let 0 = k ∈ (C) and Ey be a proper closed subspace of Ex such
that (A − kI)Ex ⊆ Ey . Then there is some x ∈ Ex such that ||x|| = 1 and
for all y ∈ Ey ,
|k|
||Ax − Ay|| ≥ .
2
(b) Let A be a compact linear operator mapping Ex → Ex and
k0 , k1 , . . . , be scalars with |kn | ≥ δ for some δ > 0 and n = 0, 1, 2, . . . .
|k| δ
||Ayp − Ayp+1 || ≥ ≥ , p = 0, 1, 2, . . . .
2 2
If follows that {yp } is a bounded sequence in Ex and
δ
||Ayp − Ayr || ≥ , p, r = 0, 1, with p = r.
2
The above shows that {Ayp } cannot have a convergent subsequence. But
this contradicts the fact that A is compact. Hence there is some nonnegative
integer p such that Ep = Ep+1 .
It can similarly be proved that there is some nonnegative integer q such
that E q+1 = E q .
8.2.4 Definitions: ρ(A), δ(A), σe (A), σa (A)
In view of the discussion in 4.7.17–4.7.20, we write the following
definitions:
4+
(i) Resolvent set: ρ(A) : {λ ∈ ( ) : A − λI is invertible}.
4+
(ii) Spectrum σ(A) : {λ ∈ ( ) : A − λI does not have an
inverse}. A scalar belonging to σ(A) is known as spectral value of A.
4 +
(iii) Eigenspectrum σe (A) of A consists of all λ in ( ), such that
A is not injective or one-to-one. Thus, λ ∈ σe (A) if and only if there is some
non-zero x in Ex such that Ax = λx. λ is called an eigenvalue of A and x
is called the corresponding eigenvector of A. The subspace N (A − λI) is
known as the eigenspace of A, corresponding to the eigenvalue λ.
(iv) The approximate eigenspectrum σa (A) consists of all λ in
4 + ( ), such that (A − λI) is not bounded below. Thus, λ ∈ σa (A) if
and only if, there is a sequence in Ex such that ||xn || = 1 for each n
and ||Axn − λxn || → 0 as n → ∞. Then λ is called an approximate
292 A First Course in Functional Analysis
If we replace A with Pn (A) and λ with λn and follow the arguments put
forward above, we conclude that R(Pn (A) − λn I) = En is a closed subspace
of Ex .
Since En+1 ⊆ En and En+1 = (A − λI)(En ) and part (b) of lemma
8.2.3 shows that that there is a non-negative integer p with Ep+1 = Ep . If
p = 0 then E1 = E0 . If p > 0, we want to show that Ep = Ep−1 .
Let y ∈ Ep−1 , that is, y = (A − λI)p−1 x for some x ∈ Ex . Then
(A − λI)y = (A − λI)p x ∈ Ep = Ep+1 , so that there is some x ∈ Ex
with (A − λI)y = (A − λI)p+1 x. Since (A − λI)(y − (A − λI)p x) = θ
and since (A − λI) is one-to-one, it follows that y − (A − λI)p x = θ, i.e.,
y = (A − λI)p x ∈ Ep . Thus, Ep = Ep−1 . Proceeding as in above, if
p > 1, we see that Ep+1 = Ep = Ep−1 = Ep−2 = · · · = E1 = E0 . But
E1 = R(A − λI) and E0 = Ex . Hence A − λI is one-to-one.
Being bounded below and onto, (A − λI) has an inverse. Hence, every
non-zero spectral value of A is an eigenvalue of A. Since σe (A) ⊂ σ(A)
always, the proof (a) is complete.
(b) Let Ex be infinite dimensional. Let us consider an infinite linearly
independent set {e1 , e2 , . . .} of Ex and let E n = span {e1 , e2 , . . . en }, n =
1, 2, . . .. Then E n is a proper subspace of E n+1 . E n is of finite dimension
and is closed by theorem 2.3.4. By the Riesz lemma (theorem 2.3.7), there
is some element an+1 ∈ E n+1 such that ||an+1 || = 1, dist (an+1 , E n ) 12 .
Let us assume that A is bounded below i.e., ||Ax|| ≥ m||x|| for all x ∈ Ex
and some m > 0. Then for all p, q = 1, 2, . . ., and p = q, we have,
m
||Aap − Aaq || ≥ m||ap − aq || ≥ ,
2
so that {Aap } cannot have a convergent subsequence, which contradicts the
fact that A is compact.
Hence, A is not bounded below. Hence 0 ∈ σa (A).
(c) If Ex is finite dimensional and D(A) = Ex , then the operator A
can be represented by a matrix, (aij ); then A − λI is also represented by a
matrix and σ(A) is composed of those scalars λ which are the roots of the
equation
a11 − λ a12 a1n
= 0 [see Taylor, [55]]
a a a −λ
n1 n2 nn
n
λn+1 xn+1 = A(xn+1 ) = αi Axi = α1 λ1 x1 + α2 λ2 x2 + · · · + αn λn xn .
i=1
4+
Proof: Let λ ∈ ( ) be such that (A − λI) is invertible, i.e., (A − λI)
has a bounded inverse. If (A − λI)B = I = B(A − λI) for some bounded
linear operator B mapping Ex → Ex , then by 6.1.5(ii) B ∗ (A∗ − λI) = I =
(A∗ − λI)B ∗ , where A∗ , B ∗ stand for adjoints of A and B respectively.
Hence σ(A∗ ) ⊆ σ(A).
Let Ex be a Banach space. By 8.2.4(iii) λ ∈ σ(A) if and only if either
A − λI is not bounded below or R(A − λI) is not dense in Ex . As because
(A − λI) is not bounded below λ ∈ σa (A).
Let f ∈ Ex∗ . Then (A∗ − λI)f = 0 if and only if f ((A − λI)x) =
∗
(A − λI)f (x) = 0 for every x ∈ Ex .
Now, (A∗ − λI) is one-to-one, i.e., N (A∗ − λI) = {θ} if and only if f = θ
wherever f (y) = 0 for every y ∈ R(A − λI). This happens if and only if the
closure of R(A−λI) is Ey i.e., R(A−λI) is dense in Ey . Hence λ ∈ σe (A∗ ).
Thus σ(A) = σa (A) ∪ σe (A∗ ).
296 A First Course in Functional Analysis
8.2.9 Theorem
Let Ex be a normed linear space and A be a compact operator mapping
Ex into Ex . Then
(a) dim N (A∗ − λI) = dim N (A − λI) < ∞ for 0 = λ ∈ ( ), 4+
(b) {λ : λ ∈ σe (A∗ ), λ = 0) = {λ : λ ∈ σe (A), λ = 0},
(c) σ(A∗ ) = σ(A).
Proof: (a) By theorem 8.1.12, A∗ is a compact linear operator mapping
Ex into Ex . Then, theorem 8.2.7 yields that the dimension r of N (A − λI)
and the dimension s of N (A∗ − λI) are both finite.
First we show that s ≤ r.
If r = 0, that is λ ∈ σe (A), then, by theorem 8.2.5(a), we see that
λ ∈ σ(A). Since σ(A∗ ) ⊆ σ(A) by theorem 8.2.8 we have λ ∈ σ(A∗ ). In
particular (A∗ − λI) is one-to-one, i.e., s = 0.
Next, let r > 1. Consider a basis {e1 , e2 , . . . , er } of N (A − λI). Then
from 4.8.3 we can find f1 , . . . , fr in Ex∗ such that fj (ei ) = δi,j , i, j =
1, 2, . . . , r.
Let, if possible, {φ1 , φ2 , . . . , φr+1 } be a linearly independent subset of
N (A∗ − λI) containing (r + 1) elements. By 4.8.2 there are y1 , y2 , . . . , yr+1
in Ex such that
φj (yi ) = δij , i, j = 1, 2, . . . , r + 1.
Consider the map B : Ex → Ex given by
r
B(x) = fi (x)yi , x ∈ Ex .
i=1
φj (A − B − λI)(x) = (A∗ − λI)(φj )(x) − φj fi (x)yi
i=1
r
=0− fi (x)φj (yi ).
i=1
−fi (x) if 1 ≤ j ≤ r
= .
0 if j = r + 1
0 = fj (x) = fj (α1 e1 + α2 e2 + · · · + αr er ) = αj , j = 1, 2, . . . , r
so that x = 0 · e1 + 0 · e2 + · · · + 0 · er = θ.
Thus A − B − λI is one-to-one because
(A − B − λI)x = θ ⇒ x = θ.
{λ : λ ∈ σ(A), λ = 0} = {λ : λ ∈ σe (A), λ = 0}
{λ : λ ∈ σ(A∗ ), λ = 0} = {λ : λ ∈ σe (A∗ ), λ = 0}
It follows from (b) above that
{λ : λ ∈ σ(A∗ ), λ = 0} = {λ : λ ∈ σ(A), λ = 0}
298 A First Course in Functional Analysis
8.2.10 Examples
,
ξ1 ξ 2 ξ 3
1. Let Ex = lp , 1 ≤ p ≤ ∞ and Ax = , , ··· where
1 2 3
x = {ξ1 , ξ2 , ξ3 , . . .} ∈ lp .
,
1 1 1
Let An = , , · · · , , 0 · · · 0 . Since An is finite, An is a linear
1 2 n
compact operator [see theorem 8.1.13].
∞ ∞
1
Furthermore, ||(A − An )x||pp = |ηi |p = |ξ |p
p i
i=n+1 i=n+1
i
∞
1 ||x||p
|ξi | p
, p > 1.
(n + 1)p i=n+1 (n + 1)p
||(A − An )x|| 1
Hence ||(A − An )|| = sup ≤ , p > 1.
||x|| n+1
x4 x5 T
Ax = x3 , , , · · · for x = (x1 , x2 , . . .)T ∈ lp
3 4
⎛ ⎞
x1 ⎛ ⎞
⎛ ⎞ x3
0 0 1 0 ··· ⎜ x2 ⎟
⎜ ⎟ ⎜ x34 ⎟
⎜ 0 0 0 1 ⎟
··· ⎠⎜ x3 ⎟ ⎜ ⎟
i.e., ⎝ 3 ⎜ ⎟ = ⎜ x5 ⎟
1 ⎜ x4 ⎟ ⎝ 4 ⎠
0 0 0 4 ··· ⎝ ⎠ ..
.. .
.
Compact Operators on Normed Linear Spaces 299
1 1
Hence A∗ can be identified with B on lp , + = 1, so that
p q
⎛ ⎞
0 0 0···
⎜ 0 0 0··· ⎟
⎜ ⎟
⎜ 0··· ⎟
B=⎜ 1 0 ⎟
⎜ 0 1
0··· ⎟
⎝ 3 ⎠
.. .. 1
. . 4
···
x2 x3 T
Hence Bx = 0, 0, x1 , , ··· for x = (x1 , x2 , x3 , . . .) ∈ lq .
⎛ ⎞ 3⎛ 4 ⎞
1 1
⎜ 0 ⎟ ⎜ 0 ⎟
⎜ ⎟ ⎜ ⎟
Since A ⎜ . ⎟ = 0 ⎜ . ⎟ we see that 0 is an eigenvalue of A. But
⎝ .. ⎠ ⎝ .. ⎠
0 0
since B is one-to-one, 0 is not an eigenvalue of B. Also since B ∗ = A, we
see that not only the compact operator B does not have an eigenvalue 0,
its adjoint B ∗ does not have an eigenvalue 0 too.
(c) Let Ex = C([0, 1]), 1 ≤ p ≤ ∞. For x ∈ Ex , let
s 1
Ax(s) = (1 − s) tx(t)dx(t) + s (1 − t)x(t)dx(t), s ∈ [0, 1] (8.1)
0 s
s 1
λx (s) = (1 − s)sx(s) − tx(t)dx(t) − s(1 − s)x(s) + (1 − t)x(t)dx(t)
0 s
s 1
=− tx(t)dx(t) + (1 − t)x(t)dx(t).
0 s
This shows that x is a continuously differentiable function, and for all
s ∈ [0, 1], we have,
λx (s) = −sx(s) − (1 − s)x(s) = −x(s).
Au − u = v (8.4)
or, Pu = v (8.5)
where P = A − I. Together with equation (8.5), consider
A∗ f − f = g (8.6)
∗
or, P f =g (8.7)
where A∗ is the adjoint operator of A and acts into the space E ∗ . By
theorem 8.1.12, A∗ is a compact operator.
8.3.2 Lemma
Let N be a subspace of the null space of the operator P , that is, a
collection of elements u such that P u = θ. Then N is a finite-dimensional
subspace of E.
Proof: Let M be an arbitrary bounded set in N . For every u ∈ N, Au = u,
that is, the operator A leaves the element of the subspace N invariant and
in particular, carries the set M into itself. The subspace N of E is then
said to be invariant with respect to A.
As A is a compact operator, A carries M into a compact set.
Consequently, every bounded set M ⊆ N is compact, implying by theorem
2.3 that N is finite dimensional.
8.3.3 Remark
The elements of the subspace N are eigenvectors of the operator A
corresponding to the eigenvalue λ0 = 1. The above conclusion remains
valid if λ0 is replaced by any non-zero eigenvalue.
Thus a compact linear operator can have only a finite number of linearly
independent eigenvectors corresponding to the same non-zero eigenvalue.
8.3.4 Lemma
Let L = P (E), that is, L be a collection of elements v ∈ E representable
in the form Au − u = v. Then L is a subspace.
To prove L is linear we note that if Au1 − u1 = v1 and Au2 − u2 = v2 ,
4
then α1 v1 + α2 v2 = A(α1 u1 + α2 u2 ) − (α1 u1 + α2 u2 ), α1 , α2 ∈ (C). Thus,
v1 , v2 ∈ L ⇒ α1 v1 + α2 v2 ∈ L. We next prove that L is closed. We first
show that there is a constant m depending only on A−I such that wherever
the equation P u = v is solvable, at least one of the solutions satisfies the
inequality
m||u|| ≤ ||v||, m > 0 (8.8)
The sequence {||u0 + wn ||} has a limit and is hence bounded. However, the
sequence {||wn ||} is also bounded, since,
F (w0 ) = ||u0 + w0 || = d.
un → u
A 0 . (8.12)
n = A
However, since u un − vn ,
n → u
u o since vn → θ.
and consequently,
un → A
A u0 (8.13)
From (8.12) and (8.13) it follows that
0 , that is, u
u0 = u
A 0 ∈ N.
Thus ||
u||/||v|| is bounded and if m = inf {||v||/||
u||}, the inequality (8.8)
is proved.
Now, suppose we are given a sequence vn ∈ L convergent to v0 . We can
assume that for some subsequence
1 1
||vnp +1 − vnp || < , where ||vnp+1 − vnp || < .
2np +1 2np
1
Then m||unp || ≤ ||vnp+1 − vnp || < .
2np
∞
This estimate yields that is the sum of the
unp converges and if u
p=1
series, then
k
k
=P
Pu lim unp = lim P unp
k→∞ k→∞
p=0 p=0
k
= lim P un0 + (vnp+1 − vnp )
k
p=1
= lim vnk+1 = v0 .
k
exhibiting v0 ∈ L. Hence, L is closed.
8.3.5 Theorem
The equation (8.4) is solvable for given v ∈ E, a Banach space, if and
only if f (v) = 0 for every linear functional f , such that
A∗ f − f = θ (8.14)
f (v) = f (Au0 −u0 ) = f (Au0 )−f (u0 ) = A∗ f (u0 )−f (u0 ) = (A∗ f −f )(u0 ) = 0.
8.3.6 Remark
An equation P u = v with the property that it has a solution u if
f (v) = 0 for every f , satisfying P ∗ f = θ, is said to be normally solvable.
The essence of the theorem 8.3.5 is that
L = P (E) is closed, is a sufficient condition for P u = v to be
normally solvable.
8.3.7 Corollary
If a conjugate homogeneous equation A∗ f − f = 0 has only a trivial
solution, then the equation Au − u = v has a solution for any right-hand
side.
8.3.8 Theorem
In order that equation (8.6) be solvable for g ∈ E ∗ given, it is necessary
and sufficient that g(u) = 0 for every u ∈ E, such that
Au − u = θ. (8.15)
Au − u = v (8.4)
1
||Avp − Avq || = ||vp − (vq − T vp + T vq )|| ≥ .
2
Thus a contradiction arises from the assumption that equation (8.4) has
in the presence of a trivial solution of the equation T u = θ, a nontrivial
Compact Operators on Normed Linear Spaces 307
solution. This proves the necessary part. Next to show that the condition
is sufficient.
Suppose that the equation T u = θ has only a trivial solution. Then, by
corollary 8.3.9, the equation
A∗ f − f = g (8.6)
is solvable for any right side. Since A∗ is also a compact operator and E ∗ a
Banach space, we can apply the necessary part of the theorem just proved
to equation (8.6). Hence the equation
A∗ f − f = θ (8.14)
8.3.11 Theorem
Let us consider the pair of equations,
Au − u = θ (8.15)
and A∗ f − f = θ (8.14)
φi (uj ) = δij , i, j = 1, 2, . . . n.
n
V u = Au + φi (u)wi .
i=1
308 A First Course in Functional Analysis
n
Then fk (V u0 − u0 ) = 0, or, fk Au0 − u0 + φi (u0 )wi = 0
i=1
n
or, A∗ fk u0 − fk u0 + φi (u0 )fk (wi ) = 0
i=1
n
or, (A∗ fk − fk )u0 + φi (u0 )fk (wi ) = 0.
i=1
Since {fi } and {wi } are biorthogonal to each other, we have from the
above equation,
(A∗ fk − fk )u0 + φk (u0 ) = 0.
Since fk is a basis of the subspace of solutions of equation (8.14),
A∗ fk − fk = θ.
Hence, φk (u0 ) = 0, k = 1, 2, . . . (n < m).
Hence we have V u0 = Au0 or, Au0 − u0 = V u0 − u0 = θ.
Since u0 ∈ N and {ui } is a basis of N ,
n
u0 = ξi ui .
i=1
n
However, φj (u0 ) = ξi φj (ui ) = ξj .
i=1
Since φj (u0 ) = 0, j = 1, 2, . . . , n, ξj = 0.
Hence, u0 = θ.
Since the equation V u − u = θ has only a trivial solution, the equation
V u − u = v is solvable for any v and in particular for v = wn+1 . Let u be
a solution of this equation. Then we can write
n
fn+1 (wn+1 ) = fn+1 Au − u + φi (u ) (wi )
i=1
n
= (A∗ fn+1 − fn+1 )u + φi (u )fn+1 (wi ) = 0
i=1
m
V ∗ f = A∗ f + f (wi )φi . (8.16)
i=1
This operator is adjoint to the operator V .
It is to be shown that the equation V ∗ f − f = θ has only a trivial
solution.
For all k = 1, 2, . . . , n.
Taking note of the biorthogonality of {φi } and {ui }
m
(V ∗ f − f )uk = (A∗ f − f )uk + f (wi )φi (uk )
i=1
= f (Auk − uk ) + f (wk )
= f (wk ) (8.17)
since {uk } is one of the bases of the subspace of solutions of (8.15). Thus,
if f0 is a solution of the the equation V ∗ f − f = θ then from (8.17) it
follows that f0 (wk ) = 0, k = 1, 2, . . . , m.
Hence (8.16) yields V ∗ f0 = A∗ f0 .
Hence 0 = V ∗ f 0 − f 0 = A∗ f 0 − f 0 ,
i.e., f0 is a solution of A∗ f − f = 0.
m m
However, f0 = βi fi = f0 (wi )fi = θ,
i=1 i=1
since f0 (wi ) = 0, i = 1, 2, 3, . . ..
∗
Since V is a compact operator, by theorem (8.3.10) the equation
V ∗ f − f = g has a solution for any g, particularly for, g = φm+1 .
Therefore if f is a solution of the above equation we have
V ∗ f − f = φm+1
Therefore,φm+1 (um+1 ) = V ∗ f (um+1 ) − f (um+1 )
m
= (A∗ f − f )um+1 + f (wi )φi (um+1 )
i=1
= f (Aum+1 − um+1 ) = 0.
On the other hand, we have by construction φm+1 (um+1 ) = 1.
The contradiction obtained proves the inequality m < n to be impossible.
Thus m = n.
8.3.12 Theorem
Let us consider the equations
Au − u = v (8.4)
∗
and A f −f =g (8.5)
Au − u = θ (8.13)
A∗ f − f = θ (8.14)
have only a trivial solution or the homogeneous equations have the same
finite number of linearly independent solutions u1 , u2 , . . . , un ; f1 , f2 , . . . , fn .
In that case equation (8.4) will have a solution, if and only if,
Au − λu = v, λ = 0. (8.18)
8.3.13 Theorem
If A is a compact operator, then its spectrum consists of finite or
countable point sets. All eigenvalues are located in the interval [−||A||, ||A||]
and in the case of a countable spectrum, these have only one limit point
λ = 0.
Proof: Let us consider the operator Tλ = A − λI.
! "
Now, for λ = 0, Tλ = −λ I − λ1 A and by theorem 4.7.12, the operator
! "
I − λ1 A and hence Tλ has an inverse when |λ| 1
||A|| < 1, i.e., the
spectrum of the operator A lies on [−||A||, ||A||]. Let 0 < m < ||A||.
For a conclusive proof it will suffice to exhibit that there can exist only a
finite number of eigenvalues λ, such that |λ| ≥ m. If that be not true, it is
possible to select a sequence λ1 , λ2 , . . . , λn of distinct eigenvalues, and also
|λi | ≥ m. Let u1 , u2 , . . . , un be a sequence of eigenvectors corresponding to
these eigenvalues, such that
Aun = λn un .
It is required to show that the elements u1 , u2 , . . . , uk for every k are linearly
independent. For k = 1, this is trivial. Suppose that u1 , u2 , . . . , uk are
linearly independent.
u − Au = v and 0 = v0 .
u0 − Au
Also, −1 v0
u − u0 = (I − A)−1 v − (I − A)
−1 ]v + (I − A)
= [(I − A)−1 − (I − A) −1 (v − v0 ).
||(I − A)−1 || ||(I − A)−1 ||
Hence, ||u − u0 || ≤ ||v|| + ||v − v0 ||.
1−
1−
Compact Operators on Normed Linear Spaces 313
8.4.2 Theorem
Let A be an operator of finite rank on a normed linear space E over
4+( ) given by
= f1 (u)u1 + · · · + fm (u)um , u ∈ E
Au
4+
(b) Let 0 = λ ∈ ( ). Them λ is an eigenvalue of A if and only if λ is
an eigenvalue of M .
Furthermore, if u (resp., u0 ) is an eigenvector of M (resp., A)
corresponding to λ, then
1 u1 + u
u0 = u 2 u2 + · · · + u
m um .
= (f1 (u0 ), . . . , fm (u0 ))T is an eigenvector of A
(resp., u (resp., M )
corresponding to λ,
0 = v0 and u
Proof: Let u0 − Au = (f1 (u0 ), . . . , fm (u0 ))T
Then for i = 1, 2, . . . , m,
)(i) = fi (u1 )f1 (u0 ) + · · · + fi (um )fm (u0 )
(M u
= fi (f1 (u0 )u1 + · · · + fm (u0 )um )
0 ) = fi (u0 − v0 ) = fi (u0 ) − fi (v0 )
= fi (Au
(i) − v(i) .
=u
− Mu
Hence u = v.
314 A First Course in Functional Analysis
1 u1 + · · · + u
=u m um = u0 − v0 .
Also for i = 1, 2, . . . , m
(i) = v(i) + (M u
u )(i) = v(i) + fi (u1 )
u(1) + fi (u2 )
u(2)
u(m)
+ · · · + fi (um )
= v(i) + fi (
u(1) u1 + u
(2) u2 + · · · + u
(m) um )
= v(i) + fi (u0 − v0 ) = fi (u0 ),
i.e., = (f1 (u0 ), . . . , fm (u0 ))T .
u
0=θ
u0 − μAu and (1) u1 + · · · + u
u0 = u (m) um .
(1) u1 + u
u0 = u (2) u2 + · · · + u
(m) um .
8.4.3 Theorem
Let E be a Banach space and A be a compact operator on E. For
n = 1, 2, . . ., let Pn ∈ (E → E) be a projection of finite rank and
AP S G
n = Pn A, An = APn , An = Pn APn .
8.4.4 Remark
Definitions AP S G
n , An , An
(ii) ASn = APn is called the Sloan projection in the name of the
mathematician Sloan.
(iii) AG
n = Pn APn is called the Galerkin projection in the name of the
mathematician Galerkin.
Now each fk ∈ E and hence each Pn ∈ (E → E). Now, Pn2 = Pn and each
∗
Pn is of finite rank.
The very definition of the Schauder basis implies that Pn u → u in E
for every u ∈ E. Hence ||A − AP n || → 0 if A is a compact operator mapping
E → E.
2. Projection of an element in a Hilbert space
Let H be a separable Hilbert space and {u1 , u2 , . . .} be an orthonormal
basis for H [see 3.8.8]. Then for n = 1, 2, . . .,
n
Pn u = < u, uk > uk , u ∈ H,
k=1
where < · > is the inner product on H. Note that each Pn is obtained by
truncating the Fourier expansion of u ∈ H [see 3.8.6]. Since H ∗ can be
identified with H (Note 5.6.1) and Pn∗ can be identified with Pn we obtain
||A−ASn || → 0 and ||A−AG n || → 0, in addition to ||A−An || → 0 as n → ∞.
P
n
(n) (n)
Pn (x) = x(tj )uj .
j=1
j=1
n
(n) (n)
≤ |x(tj ) − x(t)|uj (t)
j=1
n
(n)
≤
uj (t) =
.
j=1
Thus the kernel can be expressed as the sum of the products of functions,
one depending exclusively on t and the other depending exclusively on s. ai
and bi belong to E, i = 1, 2, . . . , m.
318 A First Course in Functional Analysis
8.4.7 Examples
Let us consider the infinite dimensional homogeneous system of
equations
∞
xi − ai,j xj = 0, i = 1, 2, . . . ∞. (8.23)
j=1
⎡ ⎤
y1
⎢ y2 ⎥
⎢ ⎥
Yn = ⎢ .. ⎥
⎣ . ⎦
yn
Since the homogeneous equation (8.23) has the only solution as zero,
(i)
the equation (8.25) has a unique solution. If we let Xn = 0, for
i = n+1, . . . then, the sequence {Xn } converges in l2 to the unique solution
⎛ ⎞
x̂1
⎜ x̂2 ⎟
⎜ . ⎟
X̂ = ⎜ . ⎟
⎜ . ⎟ of the denumerable system given by
⎝ x̂n ⎠
..
.
∞
xi − aij xj = yi , i = 1, 2, . . . ∞. (8.27)
j=1
||(I − A)−1 ||
In fact, ||X − Xn || ≤ (
n ||Y ||2 + ||Y − Yn ||2 )
1 −
n
1
provided
n = ||A − An || < ,
||(I − A)−1 ||
⎛ ⎞
∞
where AX = ⎝ aij xj ⎠ , X ∈ l2 , i = 1, 2, . . . , n, . . . , ∞
j=1
⎛⎧ n
⎪
⎨
⎜ aij xj if i = 1, 2, . . . , n
An X = ⎝ j=1
⎪
⎩
0 if i > n
These results follow from theorem 8.4.1 and theorem 8.4.3 if we note
that A is a compact operator and if
Pn X = (x1 , . . . , xn , 0, . . . , 0)T , X ∈ E,
X − An X = (y1 , . . . , yn , 0, 0, . . .)T , X ∈ l2 .
1 1 < f, uj >
u= f+ λj + g,
λ λ λ − λj
λj=λ
9. Let Ex = C([0, 1]), k(·, ·) ∈ C([0, 1] × [0, 1]) and P be the Fredholm
integral operator with kernel k ( , ).
(a) If |μ| < 1/||k||∞ , then for every v ∈ C([0, 1]), show that there is a
unique u ∈ C([0, 1]) such that u − μP u = v. Further, let
n 1
un (s) = v(s) + μj k (j) (s, t)v(t)dσ(t) , s ∈ [0, 1],
j=1 0
|μ|n+1 ||k||n+1
∞
||un − u|| ≤ ||v||∞
1 − |μ| ||k||∞
ELEMENTS OF
SPECTRAL THEORY
OF SELF-ADJOINT
OPERATORS IN
HILBERT SPACES
323
324 A First Course in Functional Analysis
It can be easily seen that the definition of adjoint operator derived here
formally coincides with the definition given in 5.6.14, for the case of Banach
spaces. It can be easily proved that the theorems on adjoint operators for
Banach spaces developed in 5.6 remain valid in complex Hilbert space too.
9.1.1 Lemma
Given a complex Hilbert space H, the operator A∗ adjoint to a bounded
linear operator A is bounded and ||A|| = ||A∗ ||.
Proof: ||Ax||2 = Ax, Ax = x, A∗ Ax ≤ ||x|| ||A∗ Ax||
||Ax|| ||A∗ Ax||
Or, sup ≤ sup . Hence, ||A|| ≤ ||A∗ ||.
x=θ ||x|| Ax=θ ||Ax||
Similarly, considering ||A∗ x||2 we can show that ||A∗ || ≤ ||A||.
Hence, ||A|| = ||A∗ ||, showing that A∗ is bounded.
9.1.2 Lemma
In H, A∗∗ = A.
Proof: The operator adjoint ot A∗ is denoted by A∗∗ .
We have A∗ x, y = y, A∗ x = Ay, x = x, Ay, ∀ x, y ∈ H
Therefore, A∗ x, y = x, A∗∗ y = x, Ay showing that A∗∗ = A.
Elements of Spectral Theory of Self-Adjoint Operators. . . 325
9.1.3 Remark
∗∗∗
(i) A = A∗ .
(ii) For A and B linear operators (A + B)∗ = A∗ + B ∗ .
(iii) (λA)∗ = λA∗ , λ ∈ +
(iv) (AB)∗ = B ∗ A∗ .
(v) If A has an inverse A−1 then A∗ has an inverse and (A∗ )−1 = (A−1 )∗ .
y ∈ D(A) =⇒ y ∈ D(A∗ ).
9.2.2 Examples
1. In an n-dimensional complex Euclidean space, a linear operator A can
be identified with the matrix (aij ) with complex numbers as elements. The
operator adjoint to A = (aij ) is A∗ = (aji ). A self-adjoint operator is a
Hermitian matrix if aij = aji .
If (aij ) is real then a Hermitian matrix becomes a symmetric matrix.
2. Adjoint operator corresponding to a Fredholm operator in
L2 ([0, 1]).
1
If T f = g(s) = k(s, t)f (t)dt (5.6.15), the kernel of the adjoint
0
operator T ∗ in complex L2 ([0, 1]) is k(t, s).
T is self-adjoint if k(s, t) = k(t, s).
3. In L2 ([0, 1]) let the operator A be given by Ax = tx(t) ∈ L2 ([0, 1]) with
every function x(t) ∈ L2 ([0, 1]). It can be seen that A is self-adjoint.
9.2.3 Remark
Given A, a self-adjoint operator, then
(i) λA is self-adjoint where λ is real.
(ii) (A + B) is self-adjoint if A and B are respectively self-adjoint.
(iii) AB is self-adjoint if A and B are respectively self-adjoint and AB =
BA.
(iv) If An → A in the sense of norm convergence in the space of operators
and all An are self-adjoints, then A is also self-adjoint.
Proof: The first part follows from the definition in 9.2.4. Let us
consider the bilinear hermitian form given by (9.5). Let us keep y fixed
in A(x, y) and obtain a linear functional of x. Consequently, A(x, y) =
x, y ∗ , y ∗ is a uniquely defined element. Thus we get an operator A,
defined by Ay = y∗ , and such that x, Ay = A(x, y).
Now x, A(y1 + y2 ) = A(x, y1 + y2 ) = A(x, y1 ) + A(x, y2 ) = x, Ay1 +
x, Ay2
Thus A is linear. Moreover, |x, Ay| = |A(x, y)| ≤ CA ||x|| ||y||.
Putting x = Ay, we get from the above, |Ay, Ay| ≤ CA ||Ay||
||y|| or ||Ay|| ≤ CA ||y||.
Hence ||A|| ≤ CA , showing that A is bounded. To prove self-adjointness
of A, we note that for x, y ∈ H, we have, x, Ay = A(y, x) = (y, Ax) =
Ax, y, implying that A = A∗ and A(x, y) = Ax, y.
Proof: Let m = inf Ax, x and M = sup Ax, x. The numbers m and
||x||=1 ||x||=1
M are called the greatest lower bound and the least upper bounds
respectively of the self-adjoint operator A.
Let ||x|| = 1. Then,
|Ax, x| ≤ ||Ax|| · ||x|| ≤ ||A||||x||2 = ||A||
and, consequently, CA = sup |Ax, x| ≤ ||A||. (9.6)
||x||=1
On the other hand, for every y ∈ H, we have, Ay, y ≤ CA ||y||2 .
Let z be any element in H, different from zero.
1
||Az|| 2 1
We put t= and u = Az,
||z|| t
we get ||Az||2 = Az, Az = A(tz), u
1
= {A(tz + u), (tz + u) − A(tz − u), (tz − u)}
4
1
≤ CA [||tz + u||2 + ||tz − u||2 ]
4
1
= CA [||tz||2 + ||u||2 ]
2
1 1
= CA t2 ||z||2 + 2 ||Az||2
2 t
1
= CA [||Az|| ||z|| + ||z|| ||Az||] = CA ||Az|| ||z||.
2
Hence, ||Az|| ≤ CA ||z||.
||Az||
Therefore, ||A|| = sup ≤ CA = sup (Ax, x) (9.7)
z=θ ||z|| ||x||=1
Elements of Spectral Theory of Self-Adjoint Operators. . . 329
9.4.3 Example 2
cos θ − sin θ
Let A=
sin θ cos θ
∗ cos θ sin θ
Then A =
− sin θ cos θ
∗ ∗ 1 0
Therefore, AA = A A = .
0 1
Thus A is unitary.
9.4.4 Remark
(i) If A is unitary or self-adjoint then A is normal.
(ii) The converse is not always true.
(iii) The operator A in example 9.4.2 although normal is not unitary.
(iv) The operator A in example 9.4.3 is unitary and necessarily normal.
9.4.5 Remark
If B is a normal operator and C is a bounded operator, such that
C ∗ C = I, then operator A = CBC ∗ is normal.
330 A First Course in Functional Analysis
Thus C ∗C = I but CC ∗ = I.
9.4.7 Definition
Given a linear operator A and a unitary operator U , the operator
B = U AU −1 = U AU ∗ is called an operator unitarily equivalent to
A.
9.4.8 Projection Operator
Let H be a Hilbert space and L a closed subspace of H. Then, by
orthogonal projection theorem for every x ∈ H, y ∈ L, and z ∈ L⊥ , x
can be uniquely represented by x = y + z.
Then P x = y and (I − P )x = z.
This motivates us to define the projection operator, see 3.6.2.
9.4.9 Theorem
P is a self-adjoint operator with its norm equal to one and P satisfies
P2 = P.
We first show that P is a linear operator.
Let, x1 = y1 + z1 , y1 ∈ L, z1 ∈ L⊥ and x2 = y2 + z2 , y2 ∈ L, z2 ∈ L⊥ .
Now, y1 = P x1 , y2 = P x2 .
Since αx1 + βx2 = [αy1 + βy2 ] + [αz1 + βz2 ], therefore,
P (αx1 + βy2 ) = αy1 + βy2 = αP x1 + βP x2 , α, β ∈ ( ). 4+
Hence P is linear.
Since y ⊥ z, we have ||x||2 = ||y + z||2 = y + z, y + z
= y, y + z, y + y, z + z, z
= ||y||2 + ||z||2
Thus, ||y||2 = ||P x||2 ≤ ||x||2 , i.e. ||P x|| ≤ ||x|| for every x.
Hence, ||P || ≤ 1.
Elements of Spectral Theory of Self-Adjoint Operators. . . 331
We have P x1 = y1 , P x2 = y2 .
Therefore, P x1 , x2 = y1 , P x2 = y1 , y2 .
Similarly, x1 , P x2 = P x1 , y2 = y1 , y2 .
Consequently, P x1 , x2 = x1 , P x2 .
Hence, P is self-adjoint.
Since, x = y + z, with y ∈ L and z ⊥ L, P x ∈ L for every x ∈ H.
Hence, P 2 x = P (P x) = P x for every x ∈ H.
Hence, projection P in a Hilbert space satisfies P 2 = P .
9.4.10 Theorem
Every self-adjoint operator P satisfying P 2 = P is an orthogonal
projection on some subspace L of the Hilbert space H.
Moreover, (I − P )2 = I − 2P + P 2 = I − 2P + P = I − P .
(I − P )x, y = x, (I − P )∗ y = x, y − x, P y
= x, (I − P )y for all x, y ∈ H.
332 A First Course in Functional Analysis
9.4.12 Theorem
For the projections P1 and P2 to be orthogonal, it is necessary and
sufficient that the corresponding subspace L1 and L2 are orthogonal.
Let y1 = P1 x and y2 = P2 x, x ∈ H. Let P1 be orthogonal to P2 . Then
y1 , y2 = P1 x, P2 x = x, P1 P2 x = 0, since P1 is orthogonal to P2 .
Since y1 is any element of L1 and y2 is any element of L2 , we conclude
that L1 ⊥ L2 . Similarly let L1 ⊥ L2 . Then for y, ∈ L1 and y2 ∈ L2 we
have y1 , y2 = 0 or P1 x, P2 x = x, P1 P2 x = 0 showing that P1 ⊥ P2 .
9.4.13 Lemma
The necessary and sufficient condition that the sum of two projection
operators PL1 and PL2 be a projection operator is that PL1 and PL2 must
be mutually orthogonal. In this case PL1 + PL2 = PL1 +L2 .
Hence P is a projection.
9.4.14 Lemma
The necessary and sufficient condition for the product of two projections
PL1 and PL2 to be a projection is that the projection operator i.e. PL1 PL2 =
PL2 PL1 . In this case PL1 PL2 = PL1 ∩L2 .
Furthermore, (PL1 PL2 )2 = PL1 PL2 PL1 PL2 = PL1 PL1 PL2 PL2
= PL21 PL22 = PL1 PL2 .
||PL2 x|| = ||PL2 PL1 x|| ≤ ||PL2 || ||PL1 x|| ≤ ||PL1 x|| (9.10)
||PL1 x|| ≥ ||PL2 x|| = ||x|| and since ||PL1 x|| ≤ ||x||
334 A First Course in Functional Analysis
9.4.18 Theorem
The difference P1 − P2 of two projections is a projection operator, if and
only if P2 is a part of P1 . In this case, LP1 −P2 is the orthogonal complement
of LP2 in LP1 .
(iv) A ≥ B, C ≥ D ⇒ A + C ≥ B + D.
(v) For any A, AA∗ and A∗ A are non-negative.
9.5.4 Example
Let us consider the symmetric operator B
d2 u
Bu = −
dx2
the functions u(x) being subject to the boundary conditions u(0) = u(1) =
0, the field Ω being the segment 0 < x < 1.
D(B) = {u(x) : u(x) ∈ C 2 (0, 1), u(0) = u(1) = 0}. Take H = L2 ([0, 1]).
Then, for all u, v ∈ D(B),
1 1
d2 x d2 v
Bu, v = v(x)dx = − u(x)dx = Bv, u
0 dx2 0 dx2
The above is true for n = 1. Let us suppose that (9.11) is true for n = k.
Then A2k (I − Ak )x, x = (I − Ak )Ak x, Ak x ≥ 0 since (I − Ak ) is a
positive operator. Hence A2k (I − Ak ) ≥ 0.
Analogously Ak (I − Ak )2 ≥ 0.
Hence, Ak+1 = A2k (I − Ak ) + Ak (I − Ak )2 ≥ 0
and I − Ak+1 = (I − Ak ) + A2k ≥ 0.
Consequently, (9.11) holds for n = k + 1.
Moreover, A1 = A21 + A2 = A21 + A22 + A3 = · · ·
= A21 + A22 + · · · + A2n + An+1 ,
n
whence A2k = A1 − An+1 ≤ A1 , since An+1 ≥ 0,
k=1
n
that is, Ak x, Ak x ≤ A1 x, x.
k=1
∞
Consequently, the series ||Ak x||2 converges and ||Ak x|| → 0 as
k=1
k → ∞.
n
Hence, A2k x = A1 x − An+1 x → A1 x as n → ∞ (9.12)
k=1
since B commutes with A and hence with A1 .
BA2 = B(A1 − A21 ) = BA1 − BA21
= A1 B − A21 B = (A1 − A21 )B,
i.e., B commutes with A2 .
Let B commute with Ak , k = 1, 2, . . . , n.
BAn+1 = B(An − A2n ) = An B − An BAn
= (An − A2n )B = An+1 B.
Hence B commutes with Ak , k = 1, 2, . . . , n, . . ..
n
ABx, x = ||A||BA1 x, x = ||A|| lim BA2k x, x .
n→∞
k=1
n
= ||A|| lim BAk x, Ak x ≥ 0
n→∞
k=1
Using (9.12) ABx, x = ||A||BA1 x, x = BAx, x.
9.5.6 Theorem
If {An } is a monotone increasing sequence of mutually commuting self-
adjoint operators, bounded above by a self-adjoint operator B commuting
Elements of Spectral Theory of Self-Adjoint Operators. . . 337
9.5.11 Example
Let H = L2 ([0, 1]). Let the operator A be defined by Ax(t) =
1
tx(t), x(t) ∈ L2 ([0, 1]). Then, ||Ax||2 = Ax, Ax = 0 t2 x2 (t)dt ≤ ||x||2 .
Hence, A is bounded.
1 1 1 √ √
2 2
Ax, x = tx(t)·x(t)dt = tx (t)dt = ( tx(t))( tx(t))dt = (Bx, Bx)
0 0 0
√
where Bx(t) = + tx(t).
Problems
1. Suppose A is linear and maps a complex Hilbert space H into itself.
Then, if Ax2 , x ≥ 0andAx, x = 0 for each x ∈ H, show that A = 0.
2. Let {u1 , u2 , . . .} be an orthonormal system in a Hilbert space H, T ∈
(H → H) and ai,j = T uj , ui , i, j = 1, 2, . . . Then show that the
matrix {ai,j } defines a bounded linear operator Q on H with respect
to u1 , u2 , u3 · · · . Show further that Q = P T P , where
Px = x, uj uj , x ∈ H.
j
du
u= = 0 at x = 0 and x = L
dx
Show that the operator A is symmetric on its domain.
8. For x ∈ L2 [0, ∞[, consider
; ∞
2 d sin us
U1 (x)u = x(s)dμ(s) [see ch. 10]
π du 0 s
; ∞
2 d 1 − cos us
U2 (x)u = x(s)dμ(s)
π du 0 s
Elements of Spectral Theory of Self-Adjoint Operators. . . 341
Show that
(i) U1 (x)u and U2 (x)u are well-defined for almost all u ∈ [0, ∞)
(ii) U1 (x), U2 (x) ∈ L2 [0, ∞[
(iii) The mappings U1 and U2 are bounded operators on L2 [0, ∞[,
which are self-adjoint and unitary.
9. Let A ∈ (H → H) be self-adjoint. Then show that
(i) A2 ≥ 0 and A ≤ ||A||I.
(ii) if A2 ≤ A then 0 ≤ A ≤ I.
Then, Ax = λx.
Pre-multiplying both sides with x∗
we have x∗ Ax = λx∗ x. (9.18)
Taking adjoint of both sides we have,
x∗ A∗ x = λx∗ x. (9.19)
From (9.18) and (9.19) it follows that
x∗ Ax = λx∗ x = λx∗ x, showing that λ = λ i.e. λ is real.
9.6.2 Lemma
Eigenvectors belonging to different eigenvalues of a self-adjoint operator
4+
in a Hilbert space H over ( ) are orthogonal.
Let x1 , x2 be two eigenvectors of a self-adjoint operator corresponding
to different eigenvalues λ1 and λ2 .
342 A First Course in Functional Analysis
Then we have
Ax1 = λ1 x1 , (9.20)
Ax2 = λx2 (9.21)
Premultiplying (9.20) by x∗2 and (9.21) with x∗1 we have,
9.6.3 Theorem
For the point λ to be a regular value of the self-adjoint operator A, it is
necessary and sufficient that there is a positive constants C, such that
Ax0 − λx0 , x = 0.
1
i.e. Rλ is a bounded operator and ||Rλ || ≤ .
C
9.6.4 Corollary
The point λ belongs to the spectrum of a self-adjoint operator A if and
only if there exists a sequence {xn } such that
9.6.5 Theorem
The spectrum of a self-adjoint operator A lies entirely on a segment
[m, M ] of the real axis, where
On the other hand, for every y ∈ H it follows from Lemma 9.3.2 that
Ay, y ≤ CA ||y||2 .
344 A First Course in Functional Analysis
Ax, x Ax, x
Let m = inf , M = sup (9.26)
x=θ ||x||2 x=θ ||x||
2
Ax, x
i.e., m≤ ≤M
||x||2
Let λ1 be any eigenvalue of A and x1 the corresponding eigenvector.
Then Ax1 = λ1 x1 and m ≤ λ1 ≤ M .
9.6.6 Theorem
M and m belong to the point spectrum.
Proof: If A is replaced by Aμ = A − μI, then the spectrum is shifted
by μ to the left and M and m change to M − μ and m − μ respectively.
Thus without loss of generality it can be assumed that 0 ≤ m ≤ M . Then
M = ||A|| [see lemma 9.3.2].
We next want to show that M is in the point spectrum. Since M = ||A||,
we can consider a sequence {xn } with ||xn || = 1 such that
Axn , xn = M −
n ,
n → 0 as n → ∞.
Further, ||Axn || ≤ ||A|| ||xn || = ||A|| = M .
Therefore, ||Axn − M xn ||2 = Axn − M xn , Axn − M xn
= Axn , Axn − 2M Axn , xn + M 2 ||xn ||2 .
= ||Axn ||2 − 2M (M −
n ) + M 2 ≤ M 2 − 2M (M −
n ) + M 2 .
= 2M
n .
Elements of Spectral Theory of Self-Adjoint Operators. . . 345
√
Hence, ||Axn − M xn || = 2M
n .
Therefore, ||Axm − M xn || → 0 as n → ∞ and ||xn || = 1.
Using corollary 9.6.4, we can conclude from the above that M belongs to
the spectrum. Similarly, we can prove that m belongs to the spectrum.
9.6.7 Examples
1. If A is the identity operator I, then the spectrum consists of
the single
eigenvalue
1 for which the corresponding eigenspace H1 = H.
1
Rλ = I is a bounded operator for λ = 1.
(λ − 1)
2. The operator A : L2 ([0, 1]) → L2 ([0, 1]) is defined by Ax = tx(t), 0 ≤
t ≤ 1.
Example 9.5.11 shows that A is a non-negative operator. Here m = 0
and M ≤ 1. Let us show that all the points of the segment [0, 1] belong to
the spectrum of A, implying that M = 1.
Let 0 ≤ λ ≤ 1 and
> 0. Let us consider the interval [λ, λ +
] or
[λ −
, λ] lying in [0, t].
⎧
⎨ √1 for t ∈ [λ, λ +
]
x (t) =
⎩ 0 for t ∈ [λ, λ +
]
1 λ+
1
Since x2 (t)dt = dt = 1.
0 λ
Hence, x (t) ∈ L2 ([0, 1]), ||x || = 1.
Furthermore, Aλ x (t) = (t − λ)x (t).
2 1 λ+
2
Therefore, ||Aλ x (t)|| = (t − λ)2 dt = .
λ 3
9.7.2 Example
Let λ be the eigenvalue of A, and Nλ the collection of eigenvectors
corresponding to this eigenvalue which includes zero as well.
Since Ax = λx, x ∈ Nλ ⇒ Ax ∈ Nλ . Hence Nλ is an invariant
subspace.
9.7.3 Remark
If the subspace L is invariant under A, we say that L reduces the
operator A.
9.7.4 Lemma
For self-adjoint A, the invariance of L implies the invariance of its
orthogonal complements, M = H − L.
Let x ∈ M , implying x, y = 0 for every y ∈ L. However, Ay ∈ L for
y ∈ L, and x, Ay = 0, i.e., Ax, y = 0 for every y ∈ L. Hence x ⊥ L and
Ax ⊥ L implies M is invariant under A. Moreover, M = H − L. Let Gλ
denote the range of the operator Aλ , i.e., the collection of all elements of
the form y = Ax − λx, λ an eigenvalue. We want to show that
H = Gλ + Nλ . Let y ∈ Gλ , u ∈ Nλ ,
then y, u = Ax − λx, u = x, Au − λu = x, 0 = 0.
Consequently, Gλ ⊥ Nλ . If y ∈ Gλ and y ∈ Gλ ,
then y = lim yn , where yn ∈ Gλ · yn , u = 0
n
⇒ y, u = limn yn , u = 0.
Consequently, Gλ ⊥ Nλ .
Now, let y, u = 0 for every y ∈ Gλ . For any x ∈ H,
0 = Ax − λx, u = x, Au − λu ⇒ Au = λu
since x is arbitrary. Therefore, u ∈ Nλ .
Consequently, Nλ = H − Gλ = H − Gλ .
9.7.5 Lemma
Gλ is an invariant subspace under a self-adjoint operator A where Gλ
stands for the range of the operator Aλ .
Proof: Let N denote the orthogonal sum of all the subspaces Nλ , i.e., a
closed linear span of all the eigenvectors of the operator A. If H is separable,
then it is possible to construct in every Nλ a finite or countably orthonormal
system of eigenvectors which span Nλ for a particular λ. Since the
eigenvectors of distinct members of Nλ are orthogonal, by combining these
systems, we obtain an orthogonal system of eigenvectors {en }, contained
completely in the span N .
The operator A defines in the invariant subspace L an operator AL in
Elements of Spectral Theory of Self-Adjoint Operators. . . 347
1
⇒ |λ| ≤ ||An || n for λ ∈ σ(A).
1
Hence, rσ (A) = Sup {|λ| : λ ∈ σ(A) ≤ ||An || n }.
1
This gives rσ (A) ≤ lim inf ||An || n (9.29)
n→∞
Further, in view of theorems 4.7.21, the resolvent operator is represented
by
Rλ (A) = −λ−1 λ−k Ak , |λ| ≥ ||A||.
Also, we have Ax = λx where λ is an eigenvalue and x the corresponding
eigenvector. Therefore
|λ| ||x|| = ||Ax|| ≤ ||A|| ||x||
or, |λ| ≤ ||A|| for any eigenvalue λ.
Hence, rσ (A) ≤ ||A||. Also Rσ (A) is analytic at every point λ ∈ σ(A). Let
x ∈ Ex and f ∈ Ex∗ . Then the function
∞
g(λ) = f (Rλ (A)x) = −λ−1 f (λ−n An x).
n=0
Elements of Spectral Theory of Self-Adjoint Operators. . . 349
is analytic for |λ| > rσ (A). Hence the singularitics of the function g all be
∞
in the disc {λ : |λ| ≤ rσ (A)}. Therefore, the series f (λ−n An x) forms
n=1
a bounded sequence. Since this is true for every f ∈ Ex∗ , an application
of uniform boundedness principle (theorem 4.5.6) shows that the elements
λ−n An form a bounded sequence in (Ex → Ex ). Thus,
||λ−n An || ≤ M < ∞
for some positive constant M (depending on λ).
1 1 1
Hence, ||An || n ≤ M n |λ| ⇒ lim Sup ||An || n ≤ |λ|.
n→∞
Since λ is arbitrary with |λ| ≤ rσ (A), it follows that
1
lim Sup ||An || n ≤ rσ (A). (9.30)
n→∞
1
It follows from (9.29) and (9.30) that lim ||An || n = rσ (A).
n→∞
9.8.7 Remark
The above result was proved by I. Gelfand [19].
9.8.8 Operator with a pure point spectrum
9.8.9 Theorem
Let A be a self-adjoint operator in a complex Hilbert space and let A
have a pure point spectrum.
Then the resolvent operator Rλ = (A − λI)−1 can be expressed as
1
Pn .
n
λn − λ
Proof: In this case, N = H and therefore there exists a closed orthonormal
system of eigenvectors {en }, such that
A en = λn en (9.31)
where λn is the corresponding eigenvalue.
Every x ∈ H can be written as
∞
x= cn en (9.32)
n=1
x = Ix = n Pn x or in the form I = Pn (9.35)
n
We know Pn Pm = 0, m = n (9.36)
By (9.31) and (9.35) Ax = cn Aen = λn Pn x (9.37)
n n
We can write A in the operator form. Then, (9.36) yields,
A= λn Pn (9.38)
n
Thus, Ax, x = λn cn en , cm em = λn c2n (9.39)
n n n
If λ does not belong to the closed set {λn } of eigenvalues, then there is a
d > 0 such that |λ − λn | > d.
We have Aλ x = (A − λI)x = (λn − λ)Pn x. Since Aλ has an inverse
n
and Pn commutes with A−1
λ , we have
x= (λn − λ)Pn A−1
λ x= Pn x.
n n
Premultiplying with Pm we have
Pm x = (λm − λ)Pm A−1 λ x.
1
Hence Rλ x = A−1 λ x= Pn x. (9.41)
n
λn − λ
cn
Since Pn x = cn en , Rλ x = en (9.42)
n
λn − λ
cn cn
Since λ n − λ ≤ d ,
12
1 2 ||x|| 1
||Rλ x|| ≤ cn = or ||Rλ || ≤ .
d n
d d
9.8.10 Remark
For n dimensional symmetric (hermitian) matrices we have similar
expressions for Rλ , with the only difference that for n-dimensional matrices
the sum is finite.
Hilbert demonstrated that the class of operators with a pure point
spectrum is the class of compact operators.
Problems
∞
f, φn
u0 = φn .
n=1
λn
MEASURE AND
INTEGRATION IN Lp
SPACES
(X, ρ) is a metric space. But it is not complete [see example in note 1.4.11].
⎧
⎪ 1
⎪
⎪ 0 if a ≤ t ≤ c −
⎪
⎨ n
Let xn (t) = 1
⎪
⎪ nt − nc + 1 if c − ≤ t ≤ c
⎪
⎪ n
⎩
1 if c ≤ t ≤ b
354
Measure and Integration in LP Spaces 355
k
Thus, fn (t) = 1 if t = , k = 0, 1, 2, . . . , n!
n!
= 0 otherwise
1
Define, ρ(fn , fm ) = |fn (t) − fm (t)|dt.
0
Now, for any value of n, ∃ some common point at which the functional
values of fn , fm take the value 1 and hence their difference is zero. On
the other hand, at the remaining points {(n! + 1) − (m! + 1)} the value
fn (t) − fm (t) is equal to 1 or −1. But the number of such types of functions
is finite and hence fn − fm = 0 only for a finite number of points. Thus
ρ(fn , fm ) = 0.
Hence, {fn } is a Cauchy sequence. Hence, {fn (t)} tends to a function
f (t) s.t.
f (t) = 1, at all rational points in 0 ≤ t ≤ 1
(10.3)
= 0, irrational points in (0, 1)
Therefore,
1 if we consider the integration in the Riemann sense, the
integral |f (t)|dt does not exist. Hence the space is not complete. We
0
would show later that if the integration is taken in the Lebesgue sense, then
the integral exists.
3. Let us define a sequence {Ωn } of sets as follows:
Ω0 = [0, 1]
1
Ω1 = Ω0 with middle open interval of length 4 removed.
Ω2 = Ω1 with middle open interval of the component intervals of Ω1
removed, each of length 412 .
Then by induction we have already defined Ωn so as to consist of 2n
disjoint closed intervals of equal length. Let Ωn+1 = Ωn with middle open
1
intervals of the component intervals of Ωn removed, each of length 4n+1 .
356 A First Course in Functional Analysis
10.1.2 Remark
In example 2 (10.1.1) it may seen that {fn } → 1 at all rational points
in [0, 1] and −→ 0 at all irrational points in [0, 1].
It is known that the set of rational points in [0, 1] can be put into one-to-
one correspondence with the set of positive integers, i.e., the set of natural
numbers [see Simmons [53]]. Hence the set of rational numbers in [0, 1]
forms a countable set. Thus, the set of rational numbers in [0, 1] can be
written as a sequence {r1 , r2 , r3 , . . .}.
Let
be any positive real number. Suppose we put an open interval
of width
about the first rational number r1 , an interval of width
/2
about r2 and so on. About rn we put an open-interval of width
/2n−1 .
Then we have an open interval of some positive width about every rational
number in [0, 1]. The sum of the widths of these open intervals is
+ 2 + 22 + · · · + 2n + · · · = 2
. We conclude from all this that all
rational numbers in [0, 1] can be covered with open intervals, the sum of
whose length is an arbitrarily small positive number.
We say that the Lebesgue measure of the R of rational numbers in
[0, 1] is l · m · (R) = 0. This means that the greatest lower bound of the
total lengths of a set of open intervals covering the rational number is zero.
The Lebesgue measure of the entire interval [0, 1] is l · m · [0, 1] = 1. This
is because the greatest lower bound of the total length of any set of open
intervals covering the whole set [0, 1] is 1.
Now if we remove the rational numbers in [0, 1] from [0, 1] we are left
with the set of irrational numbers the Lebesgue measure of which is 1.
Thus, if we delete from [0, 1] the set M of rational numbers, whose
1
Lebesgue measure is zero, we can find L f (t)dt in example 2 above, i.e.,
0
1
f (t)dt = 1 · L f (t)dt
[0,1]−M 0
denotes the integration in the Lebesgue sense. The above discussion may
be treated as a prelude to a more formal treatment ahead.
10.1.3 The Lebesgue outer measure of a set E ⊂ R
4
The Lebesgue outer measure of a set E ⊆ is denoted by m∗ (E) and
is defined as
3∞ ∞
*
m∗ (E) = g · l · b l(In ) : E ⊂ In ,
n=1 n=1
where In is an open interval in 4 and l(In) denotes the length of the interval
In .
358 A First Course in Functional Analysis
m∗ (A) ≥ m∗ (A ∩ S) + m∗ (A ∩ S C ).
10.1.6 Remark
(i) Since the definition of measurability is symmetric in S and S C , we
have S C measurable whenever S is.
(ii) Φ and the set 4 of all real numbers are measurable.
10.1.7 Lemma
∗
If m (S) = 0 then S is measurable.
Proof: Let A be any set. Then A ∩ S ⊂ S and so m∗ (A ∩ S) ≤ m∗ (S) = 0.
Also A ⊇ A ∩ S C .
Hence, m∗ (A) ≥ m∗ (A ∩ S C ) = m∗ (A ∩ S) + m∗ (A ∩ S C ).
But m∗ (A) ≤ m∗ (A ∩ S) + m∗ (A ∩ S C ).
Hence S is measurable.
Measure and Integration in LP Spaces 359
10.1.8 Lemma
If S1 and S2 are measurable, so is S1 ∪ S2 .
Proof: Let A be any set. Since S2 is measurable, we have
Since (S1 ∪S2 )C = S1C ∩S2C . Hence, S∪S2 is measurable, since the above
4
equality is valid for every set A ⊆ , where S C denotes the complement of
4
S in . If S is measurable then m∗ (S) is called the Lebesgue measure
of S and is denoted simply by m(S).
10.1.9 Remark
(i) Φ and 4 are measurable subsets.
(ii) The complements and countable union of measurable sets are
measurable.
10.1.10 Lemma
Let A be any set and S1 , S2 , . . . , Sn , a finite sequence of disjoint
measurable sets. Then
n
* n
∗
m A∩ Si = m∗ (A ∪ Si )
i=1 i=1
*
A∩ Si ∩ Sn = A ∩ Sn (10.7)
i=1
n−1
*
n *
A∩ Si ∩ SnC =A∩ Si (10.8)
i=1 i=1
* *
∗ ∗
m A∩ Si =m A∩ Si ∩ Sn
i=1 i=1
*
n
∗
+m A∩ Si ∩ SnC
i=1
360 A First Course in Functional Analysis
n−1
* *
m∗ A ∩ Si = m∗ (A ∩ Sn ) + m∗ A ∩ Si
i=1 i=1
n−1
= m∗ (A ∩ Sn ) + m∗ (A ∩ Si ) (10.9)
i=1
(i) For each real number α the set {x : f (x) > α} is measurable.
(ii) For each real number α the set {x : f (x) ≥ α} is measurable.
(iii) For each real number α the set {x : f (x) < α} is measurable.
(iv) For each real number α the set {x : f (x) ≤ α} is measurable.
If (i)–(iv) are true, then
(v) For each extended real number α the set {x : f (x) = α} is
measurable.
(ii) =⇒ (v) for α = ∞. Similarly, (iv) =⇒ (v) for α = −∞, and we have
(ii) and (iv) =⇒ (v).
10.2.4 Remark
(i) It may be noted that an extended real valued function f is (Lebesgue)
measurable if its domain is measurable and if it satisfies one of the
first four statements of the lemma 10.2.3.
(ii) A continuous function (with a measurable domain) is measurable,
because the preimage of any open set in 4
is an open set.
(iii) Each step function is measurable.
362 A First Course in Functional Analysis
10.2.5 Lemma
Let K be a constant and f1 and f2 be two measurable real-valued
functions defined on the same domain. Then the functions f1 + K, Kf1 ,
f1 + f2 , f2 − f1 and f1 f2 are measurable.
Proof: Let {x : f1 (x) + K < α} = {x : f1 (x) < α − K}.
Therefore by condition (iii) of lemma 10.2.3, since f1 is measurable,
f1 + K is measurable.
If f1 (x) + f2 (x) < α, then f1 (x) < α − f2 (x) and by the corollary
to the axiom of Archimedes [see Royden [47]] we have a rational number
between two real numbers. Hence there is a rational number p such that
f1 (x) < p < α − f2 (x).
Hence, {x : f1 (x) + f2 (x) < α} = ∪{x : f1 (x) < p} ∩ {x : f2 (x) < α − p}.
Since the rational numbers are countable, this set is measurable and so
f1 + f2 is measurable.
Since −f2 = (−1)f2 is measurable, when f2 is measurable f1 − f2 is
measurable.
√ √
Now, {x : f 2 (x) > α} = {x : f (x) > α} ∪ {x : f (x) < − α} for α > 0
and if α < 0
{x : f 2 (x) > α} = D,
where D is the domain of f . Hence f 2 (x) is measurable. Moreover,
1
f1 f2 = [(f1 + f2 )2 − f12 − f22 ].
2
Given f1 , f2 measurable functions, (f1 +f2 )2 , f12 and f22 are respectively
measurable functions.
Hence, f1 f2 is a measurable function.
10.2.6 Remark
Given f1 and f2 are measurable,
10.2.7 Theorem
Let {fn } be a sequence of measurable functions (with the same domain
of definition). Then the functions sup{f1 , . . . , fn } and inf{f1 , f2 , . . . , fn },
sup fn , inf fn , inf sup fk , sup inf fk are all measurable.
n n n k≥n n k≥n
Proof: If q is defined by q(x) = sup{f1 (x), f2 (x), . . . , fn (x)} then {x :
*∞
q(x) > α} = {x : fn (x) > α}. Since fi for each i is measurable, g is
n=1
measurable.
Measure and Integration in LP Spaces 363
n
φ(x) = ai χEi (x) (10.10)
i=1
Ei = {x ∈ 4 : φ(x) = ai}
364 A First Course in Functional Analysis
10.2.11 Remark
(i) The representation for φ is not unique.
(ii) The function φ is simple if and only if it is measurable and assumes
only a finite number of values.
(iii) The representation (10.10) is called the canonical representation and
it is characterised by the fact that Ei are disjoint and the ai distinct
and non-zero.
10.2.12 Example
Let f : 4 −→ [0, ∞[. Consider the simple function for n = 1, 2, . . .
3 (i−1) (i−1)
2n if 2n ≤ f (x) < i
2n for i = 1, 2, . . . , n2n
φn (x) =
n if f (x) ≥ n
n
when φ has the canonical representation φ = ai χEi .
i=1
10.2.14 Lemma
n
Let φ = ai χEi , with Ei ∩ Ej = ∅ for i = j. Suppose each set Ei is a
i=1
measurable set of finite measure. Then
n
φdm = ai m(Ei ).
i=1
*
Proof: The set Aa = {x : φ(x) = a} = Ei
ai =a
Measure and Integration in LP Spaces 365
Hence am(Aa ) = ai m(Ei ) by the additivity of m, and hence
ai =a
n
φ(x)dm(x) = a m(Aa ) = ai m(Ei ).
i=1
10.2.15 Theorem
Let φ and ψ be simple functions, which vanish outside a set of finite
measure. Then
(αφ + βψ)dm = α φdm + β ψdm
Proof: Let {Ei } and {Ei } be the set occurring in canonical representations
of φ and ψ. Let E0 and E0 be the sets where φ and ψ are zero. Then the
set Fk obtained by taking the intersections of Ei ∩ Ei are members of a
finite disjoint collection of measurable sets and we may write
n
n
φ= ai χFi ψ= bi χFi
i=1 i=1
and so αφ + βψ = (αai + βbi )χFi
Hence, using lemma 10.2.13
(αφ + βψ)dm = a φdm + b ψdm
and sup φ. (10.13)
φ≤f E
10.2.18 Definition
If f is a complex-valued measurable function over , then we define 4
f dm = Re f dm + i Im f dm
4 4 4
whenever Ref dm and Imf dm are well-defined.
4 4
10.2.19 Definition
If f is a measurable function on 4 and |f |dm < ∞, we say f is an
integrable function on . 4 4
In what follows we state without proof some important convergence
theorems.
10.2.20 Theorem
Let {fn } be a sequence of measurable functions on a measurable subset
E of . 4
(a) Monotone convergence theorem: If 0 ≤ f1 (x) ≤ f2 (x) ≤ · · · and
fn (x) −→ f (x) for all x ∈ E, then
fn dm −→ f dm. (10.15)
E E
Measure and Integration in LP Spaces 367
Proof: The above result is true where f1 and f2 are simple measurable
functions defined on E [see theorem 10.2.15]. We write f1 = f1+ −f1− , where
f1+ = max{f1 , 0} and f1− = min{f1 , 0}. Similarly we take f2 = f2+ − f2− ,
f2+ and f2− will have similar meanings as those of f1+ , f1− . It may be noted
that f1+ , f1− , f2+ , f2− are nonnegative functions.
We now approximate f1+ , f1− , f2+ , f2− by non-decreasing sequence
of simple measurable functions and applying the monotone convergence
theorem (10.2.20(a)).
with the infimum taken over all possible subdivisions of [a, b]. Similarly, we
define the lower integral
b
f (x)dx = sup s (10.18)
a
368 A First Course in Functional Analysis
The upper integral is always at least at large as the lower integral and
if the two are equal we say f is Riemann integrable and call the common
value, the Riemann integral of f . We shall denote it by
b
R f (x)dx (10.19)
a
for some continuous function f on [a, b]. In that case F (x) = f (x) for all
x ∈ [a, b]. For a proof see Rudin [48].
10.4.1 Absolutely continuous function
4+
A ( )-valued function F on [a, b] is said to be absolutely continuous
on [a, b] if for every
> 0, there is some δ > 0 such that
n
|F (xi ) − F (yi )| <
i=1
n
whenever a ≤ y1 < x1 < · · · < yn < xn ≤ b and (xi − yi ) < δ.
i=1
10.4.2 Remark
(i) Every absolutely continuous function is uniformly continuous on
[a, b].
(ii) If F is differentiable on [a, b] and its derivative F is bounded on
[a, b], then F is absolutely continuous by the mean value theorem.
Measure and Integration in LP Spaces 369
for some (Lebesgue) integrable function; f on [a, b]. In that case F (x) =
f (x) for almost all x ∈ [a, b] [see Royden [47]].
10.5.1 Total variation, bounded variation
4
Let f : [a, b] −→ (C) be a function. Then the (total) variation Var
(f ) of f over [a, b] is defined as
n
Var (f ) = sup |f (ti ) − f (ti−1 )| : P = [t0 , t1 , . . . , tn ]
i=1
The functions defined by these integrals are integrable on [a, b] and [c, d]
respectively.
Moreover, K(s, t)d(m × m)(s, t)
[a,b]×[c,d]
b ? d )
= K(s, t)dm(t) dm(s)
a c
d ? b )
= K(s, t)dm(s) dm(t).
c a
∞
∞
1/p n
1/q
|xi yi | ≤ |xi | p
|yi | q
By Minkowski’s inequality
p
1/p
1/p
n
|gn | dm
p
= |fi+1 − fi |
E E i=0
n
1/p
n
≤ |fi+1 − fi |p = ρp (fi+1 , fi )
i=0 E i=0
n
1
= ρp (f1 , f0 ) + i
i=1
2
n−1
We know that fn (x) = [fi+1 (x) − fi (x)] for all x ∈ E, we have
i=0
n
lim fn (x) = f (x) and |fn (x)| ≤ |fi+1 (x) − fi (x)| ≤ g(x)
n→∞
i=1
0 0
0n 0
0 0
≤ ||f || 0
i [uti (ζ) − usi (ζ)]0
0 0
i=1
p
1/p
1 n
= ||f ||
i [uti (ζ) − usi (ζ)] dζ
0
i=1
n
1/p
1/p
n
≤ ||f || dζ = ||f || m(δi )
i=1 δi i=1
Now h(0) = f [u0 (ζ)] = 0, since u0 (ζ) ≡ 0 is the null element of Lp [0, 1].
We have, t
h(t) = α(s)ds
0
n 1
If vn (s) = Ck [u k (s) − u k−1 (s)], then f (vn ) = vn (s)α(s)ds.
n n
k=1 0
Since, on the other hand, vm (t) −→ x(t) a.e. and vm (t) is uniformly
bounded, it follows that
1 1/p
||vm − x||p = |vm (t) − x(t)| dt p
−→ 0 as m → ∞
0
1 q1 1 q1
(q−p)p (q−1)p
|xn (t)| dt −→ |α(t)| dt ≤ ||f || as n → ∞
0 0
1 1/q
or, |α(t)|q ≤ ||f || i.e. α(t) ∈ Lq [0, 1] (10.22)
0
1
Now, let x(t) be any function in Lp [0, 1]. Then there exists 0 x(t)α(t)dt.
Furthermore, there exists a sequence {xm (t)} of bounded functions, such
that
1
|x(t) − xm (t)|p dt → 0 as m → ∞
0
1 1 1
Therefore, xm (t)α(t)dt − x(t)α(t)dt ≤ |(xm (t) − x(t)| |α(t)|dt
0 0 0
1 1/p 1 1q
≤ |(xm (t) − x(t)| dt · p
|α(t)| dt
q
0 0
Using the fact that xm (t) − x(t) ∈ Lp [0, 1] and α(t) ∈ Lq ([0, 1]), the
above inequality is obtained by making an appeal to Hölder’s inequality.
Since the sequence {xm (t)} are bounded and measurable functions, in
Lp ([0, 1]) and α(t) ∈ Lq ([0, 1]),
1
xm (t)α(t)dt = f (xm )
0
1
Hence, f (xm ) −→ x(t)α(t)dt as m → ∞
0
1 1/p 1 1/q
Moreover, ||g(x)||p ≤ |x(t)| dt
p
|β(t)| dt
q
< ∞,
0 0
≤ ||x||p ||α||q
Hence ||f || ≤ ||α||q (10.24)
It follows from (10.23) and (10.24) that
1 1/q
||f || = ||α(t)||q = |α(t)|q dt
0
10.7.1 Remark
(i) Kolmogoroff [29] gave an example of a function x in L1 ([−π, π]) such
that the corresponding sequence {sn (t)} diverges for each t ∈ [−π, π].
(ii) If x ∈ Lp ([−π, π]) for some p > 1, then {sn (t)} converges for almost
all t ∈ [−π, π] (see Carleson, A [10]).
UNBOUNDED LINEAR
OPERATORS
381
382 A First Course in Functional Analysis
11.2.1 Theorem
If A has a bounded inverse, then R(A∗ ) is closed.
∗
A∗ fp − A∗ fq ≥ mfp − fq .
11.2.6 Theorem
If M is a subspace of Ex∗ , then (⊥ M )⊥ ⊃ M . If Ex is reflexive, then
( M )⊥ = M .
⊥
11.2.7 Remark
If Ex is reflexive, i.e., Ex = Ex∗∗ , then (⊥ M )⊥ = M .
11.2.8 Definition: domain of A∗
Domain of A∗ is defined as
D(A∗ ) = {φ : φ ∈ Ey∗ , φA is continuous on D(A)}.
For φ ∈ D(A∗ ), let A∗ be the operator which takes φ ∈ D(A∗ ) to φA,
where φA is the unique continuous linear extension of φA to all of Ex .
11.2.9 Remark
(i) D(A∗ ) is a subspace of Ey∗ and A∗ is linear.
(ii) A∗ φ is taken to be φA rather than φA in order that R(A∗ ) is
contained in Ex∗ [see 5.6.15].
Unbounded Linear Operators 385
11.2.10 Theorem
(i) R(A)⊥ = R(A)⊥ = N (A∗ ).
(ii) R(A) = ⊥ N (A∗ ).
In particular, A has a dense range if and only if A∗ is one-to-one.
11.2.11 Theorem
If A and A∗ each has an inverse then (A−1 )∗ = (A∗ )−1 .
Hence, by induction, we can construct sequences {xk } and {fk } having the
following properties :
xk = fk = fk (xk ) = 1, Axk < , 1≤k<∞ (11.7)
3k
xk ∈ ∩k−1
i=1 N (fi ) or equivalently, fi (xk ) = 0 1 ≤ i < ∞ (11.8)
i = k
α1 x1 + α2 x2 + · · · + αk xk + · · · = 0.
or α1 fi (x1 ) + α2 fi (x2 ) + · · · + αk fi (xk ) + · · · = 0.
j
fj+1 (x) = αi fj+1 (xi ) + αj+1 (11.10)
i=1
j
|αj+1 | ≤ |fj+1 (x)| + |αi | |fj+1 (xi )|
i=1
j
≤ x + 2i−1 x = 2j x
i=1
Hence, (11.9) is true by induction.
l
l
Thus, Ax ≤ |αi |Axi ≤ 2i−1 3−i
x ≤
x
i=1 i=1
Hence, Ax ≤
.
Unbounded Linear Operators 389
n
n
AL
n x ≤ |αi | Axi ≤ 2i−1 3−i
x.
i=1 i=1
Proof: We first show that (a) implies (b). Let us suppose that B is strictly
singular and L is an infinite-dimensional subspace of Ex . Then BL , the
restriction of B to L, is strictly singular. Therefore, BL does not have a
bounded inverse on an infinite dimensional subspace M ⊆ L. Hence, by
theorem 11.3.4, B is precompact on such a M ⊆ L.
Next, we show that (b) ⇒ (c). If (b) is true then we assert that B does
not have a bounded inverse on an infinite dimensional subspace, having
finite deficiency in L. If that is not so, then B would be precompact and
390 A First Course in Functional Analysis
would have a bounded inverse at the same time. This violates the conclusion
of theorem 11.3.3. Hence by applying theorem 11.3.4 to BL (c) follows.
Finally, we show that (c) ⇒ (a). It follows from (c) that BL ≤
, i.e.,
BL x ≤
x for an arbitrary small
> 0, i.e., BL x >mx, m > 0
and finite, for all x belonging to an infinite dimensional subspace M ⊆ L.
Hence, BL does not have a bounded inverse on M ⊆ L. Thus B is strictly
singular.
Index of A: If α(A) and β(A) are not both infinite, we say A has an inverse.
The index κ(A) is defined by κ(A) = α(A) − β(A).
11.5.2 Examples
1. Let Ex = Lp ([a, b]) Ey = Lq ([a, b]) where 1 ≤ p, q < ∞.
Let us define A as follows :
D(A) = {u : u(n−1) exists and is absolutely continuous on [a, b], u(n) ∈ Ey }.
Unbounded Linear Operators 391
Au = u(n) , u(n) stands for the n-th derivative of u in [a, b]. It may
be recalled that an absolutely continuous function is differentiable almost
everywhere (10.5). Here, N (A) is the space of polynomials of degree at
most (n − 1). Hence, α(A) = n, β(A) = 0.
2. Let Ex = Ey = lp , 1 ≤ p ≤ ∞. Let {λκ } be a bounded sequence of
numbers and A be defined on all of Ex by A({xκ }) = {λκ xκ }.
α(A) are the members of λk which are 0. β(A) = 0 if {1/λκ } is a
bounded sequence. β(A) = ∞ if infinitely many of the λκ are 0.
11.5.3 Lemma
Let L and M be subspace of Ex with dim L > dim M (thus dim M <
∞). Then, there exists a l = 0 in L such that
l = dist (l, M ).
Note 11.5.1 This lemma does not hold if dim L = dim M < ∞.
4
For example, if Ex = 2 and L and M are two lines through the origin
which are not perpendicular to each other.
If Ex is a Hilbert space, the lemma has the following easy proof.
Ax
γ(A) = inf (11.11)
xD(A) d(x, N (A))
11.5.5 Definition
The one-to-one operator  of A induced by A is the operator from
D(A)/N (A) into Ey defined by
Â[x] = Ax,
392 A First Course in Functional Analysis
where the coset [x] denotes the set of elements equivalent to x and
belongs to D(A)/N (A).
 is one-to-one and linear with same range as that of A. We next state
without proof the following theorem.
11.5.6 Theorem (Goldberg [21])
Let N (A) be closed and let D(A) be dense in Ex . If γ(A) > 0, then
γ(A) = γ(A∗ ) and A∗ has a closed range.
11.5.7 Theorem
Suppose γ(A) > 0. Let V be bounded with D(V ) ⊃ D(A). If
V < γ(A), then
(a) α(A + V ) ≤ α(A)
(b) dim Ey /R(A + V ) ≤ dim Ey /R(A)
AEx is closed. But this contradicts the hypothesis that R(A) is not closed.
Therefore, A does not have a bounded inverse on U . Hence there exists
an L = L(
) with the properties described in theorem 11.3.4. Since Ey is
complete and A is closed and bounded on L, it follows that L is contained
in D(A). Moreover, AL ≤
and ABL = AB L , where BL and BL are
unit balls in L and L respectively and AL is the restriction of A to L.
The precompactness of A and the completeness of Ey imply that AB L is
compact. Thus, AL is compact.
11.6.3 Theorem
Suppose that A1 is a linear extension of A such that
Proof: (a) Since α(A) < ∞, i.e., the null space of A is finite dimensional
i.e., closed, there exists a closed subspace L of Ex such that Ex = L⊕N (A).
Let AL be the operator A restricted to L ∩ D(A). Then A being closed, AL
is closed with R(AL ) = R(A). Let us suppose that A + V does not have a
closed range. Now A + V is an extension of AL + V . Then it follows from
theorem 11.6.3(b) that AL + V does not have a closed range. Moreover,
11.6.3(a) yields that AL + V is closed since AL is closed. Thus, AL + V is a
closed operator but its range is not closed. It follows from theorem 11.6.2
that there exists a closed infinite–dimensional subspace L0 contained in
D(AL ) = D(AL + V ) such that
γ(AL )
(AL + V )x < x, x ∈ L0 (11.15)
2
Thus, since AL is one-to-one, it follows for all x in L0 ,
A1 u = f, u ∈ D(A1 ) (11.23)
A2 u = f (11.24)
11.7.5 Theorem
Let the symmetric and coercive operators A1 and A2 fulfill respectively
the inequalities (11.25) and (11.26). Moreover, let HA1 and HA2 coincide
and are each seperable. If u0 and u1 are the solutions of equations (11.23)
and (11.24), then there exists some constant η such that
both the form K(x, x) and the unit form E(x, x) are altered. Perturbation
theory is applied in this case. See Courant and Hilbert [15].
3. In what follows, we consider a differential equation where perturbation
method is used. We consider the differential equation
d2 y
+ (1 + x2 )y + 1 = 0, y(±1) = 0 (11.34)
dx2
We consider the perturbed equation
d2 y
+ (1 +
x2 )y + 1 = 0, y(±1) = 0 (11.35)
dx2
d2 y0
For
= 0, the equation + y0 + 1 = 0, y0 (±1) = 0 has the solution
dx2
cos x
y0 = − 1.
sin x
Let y(x,
) be the solution of the equation (11.33) and we expand y(x,
)
in terms of
y(x1
) = y0 (x) +
y1 (x) +
2 y2 (x) · · · (11.36)
Substituting the power series (11.36) for y in the differential equation,
we obtain,
∞ ∞
n (yn + yn ) +
n x2 yn−1 + 1 = 0
n=0 n=1
yn (±1) = 0 (11.38)
THE HAHN-BANACH
THEOREM AND
OPTIMIZATION
PROBLEMS
It was mentioned at the outset that we put emphasis both on the theory and
on its application. In this chapter, we outline some of the applications of
the Hahn-Banach theorem on optimization problems. The Hahn-Banach
theorem is the most important theorem about the structure of linear
continuous functionals on normed linear spaces. In terms of geometry, the
Hahn-Banach theorem guarantees the separation of convex sets in normed
linear spaces by hyperplanes. This separation theorem is crucial to the
investigation into the existence of an optimum of an optimization problem.
400
The Hahn-Banach Theorem and Optimization Problems 401
Xr+1 = Φ. Suppose the two sets are disjoint. Then there exists a plane P
which separates them strongly. Writing Xj = Xj ∩ P we have
*
r
Xj = ∪(Xj ∩ P ) ∪ (Xr+1 ∩ P )
j=1 ⎛ ⎞
*
r+1
= P ∩ ⎝ Xj ⎠
j=1
Therefore, the union of the sets X1 , X2 , . . . , Xr is convex. Also, the
intersection of any (r − 1) of X1 , X2 , . . . , Xr meets X and Xr+1 and hence
meets P . Therefore, the intersection of any (r − 1) of X1 , X2 , . . . , Xr is
not empty.
But by hypothesis ∩jr=1 Xj = X ∩ P = Φ contradicting the fact that P
is a hyperplane which separates X and Xr+1 strictly.
It follows that X ∩ Xr+1 = Φ and so the result holds for m = r + 1.
12.1.3 Theorem
A closed convex set is equal to the intersection of the half-spaces which
contain it.
Proof: Let X be a closed convex set and let A be the intersection of
4
the half-spaces which contains it. If Ef = {x : x ∈ n , f (x) = α} is
4 4
a hyperplane in n , then Hf = {x : x ∈ n , f (x) ≥ α} is called a
half-space.
If x0 ∈ X, then {x0 } is a compact convex set not meeting X. Therefore,
402 A First Course in Functional Analysis
||v − u0 || ≥ α ≥ φ(u0 ).
Proof of theorem: (i) and (ii) Since α is the infimum in (12.1), for each
> 0, there is a point u ∈ X such that,
||u − u0 || ≤ α + .
Hence β ≤ α +
, for all
> 0, that is, β ≤ α. Let α > 0. Now theorem
5.1.4 yields that there is a functional fˆ ∈ X ⊥ with ||fˆ|| = 1 such that
fˆ(u0 ) = α (12.6)
4
Let fˆr : X → be the restriction of fˆ0 : E → 4 to X.
Then ||fˆr || = sup fˆ0 (u) = β.
||u||≤1
u∈X
Case I Let VarJ (h) = 0 for all
> 0. Then, by (12.11), h is a step function
of the following form,
3
0 if a ≤ x < x1
h(x) =
±Var(h) if x1 < x ≤ b
406 A First Course in Functional Analysis
Since, dim X < ∞, i.e., finite, the problem (12.12) has a solution by 12.2.4.
If u0 ∈ 2, then (12.12) is immediately true. Let us assume that u0 ∈ 2.
Let p̂ be a solution of (12.12). Then, since u0 ∈ 2 and p̂ ∈ 2 we have
||u0 − p|| > 0. By the duality theory from theorem 12.2.3 there exists a
functional f ∈ C([a, b])∗ such that
Let us measure the minimal fuel expenditure during the time interval
T
[0, T ] through the integral 0 |u(t)|dt over the rocket thrust. Let T > 0
be fixed. Then the minimal fuel expenditure α(T ) during the time interval
[0, T ] is given by a solution of the following minimum problem
T
min |u(t)|dt = α(T ) (12.17)
u 0
n n
tj T
Δ= |h(tj ) − h(tj−1 )| ≤ |u(t)|dt = |u(t)|dt.
j=1 j=1 tj−1 0
T
Hence, Var(h) ≤ 0 |u(t)|dt.
By the mean value theorem,
n
Δ= |u(ξj )|(tj − tj−1 ) where tj−1 ≤ ξj ≤ tj .
j=1
and hence T
Var(h) = |u(t)|dt. (12.24)
0
VARIATIONAL
PROBLEMS
410
Variational Problems 411
φ = 0 a direction in E1 .
13.2.1 Definition
The mapping P is said to be differentiable in the sense of Gâteaux,
or simply G-differentiable at a point u ∈ U in the direction φ if the
difference quotient P (u+tφ)−P
t
(u)
has a limit P (u, φ) in E2 as t → 0 in . 4
The (unique) limit P (u, φ) is called the Gâteaux derivative of P at u
in the direction of φ.
P is said to be G-differentiable in a direction φ in a subset of U if it is
G-differentiable at every point of the subset in the direction φ.
13.2.2 Remark
The operator E1 # φ −→ P (u, φ) ∈ H is homogeneous.
P (u + tαφ) − P (u)
For P (u, αφ) = lim ·α
t→0 tα
= αP (u, φ) for α > 0.
13.2.3 Remark
The operator P (u, φ) is not, in general, linear.
4
Example 1. Let f : 2 −→ be defined by 4
⎧
⎨0 if (x1 , x2 ) = (0, 0)
f (x1 , x2 ) = x51
⎩ if (x1 , x2 ) = (0, 0)
((x1 − x2 )2 + x41 )
1 Then g is continuously
1 differentiable in ]0, 1[ and g(1) − g(0) =
g (θ)dθ = tφ(x) f (u + θtφ)dθ (θ = θ(x), |θ(x)| ≤ 1), so that
0 0
1
(J(u + tφ) − J(u))
= φ(x) f (u(x) + θtφ(x))dθdx
t Ω 0
Now, φ(x)f (u(x + θtφ(x))dx ≤ K |φ(x)||u(x + θtφ(x))|p−1 dx
Ω Ω
1/p 1/q
≤K |φ(x)|p dx |u(x + θtφ(x)|(p−1)q dx <∞
Ω Ω
13.2.4 Definition
An operator P : U (E1 −→ E2 ) (U being an open set in E1 ) is said
to be twice differentiable in the sense of Gâteaux at at point u ∈ U
in the directions φ, ψ (φ, ψ ∈ E1 , φ = 0, ψ = 0 given) if the operator
u −→ P (u, φ) : U ⊂ E1 −→ E2 is once G-differentiable at u in the
direction ψ. The G-derivative of u −→ P (u, φ) is called the second
G-derivative of P and is denoted by P (u, φ, ψ) ∈ E2 , i.e.,
P (u + tφ, ψ) − P (u, ψ)
P (u, φ, ψ) = lim (13.6)
t→0 t
13.2.5 Gradient
Let J : U ⊂ E1 −→ 4
be a functional on an open set of a normed
linear space E1 which is once G-differentiable at a point u ∈ U . If the
functional u −→ J (u, φ) is continuous linear on E1 , then there exists a
(unique) element G(u) ∈ E1∗ (6.1), such that
J (u, φ) = G(u)(φ) for all φ ∈ E1 .
Similarly, if J is twice G-differentiable at a point u ∈ U , and if the form
(φ, ψ) −→ J (u, φ, ψ) is a bilinear (bi-) continuous form on E1 × E1 , then
there exists a (unique) element H(u) ∈ (E1 −→ E1∗ ) such that
J (u, φ, ψ) = H(u)(φ, ψ), (φ, ψ) ∈ E1 × E1 (13.7)
13.2.6 Definitions: gradient, Hessian
Gradient: G(u) ∈ E1∗ is called the gradient of J at u
Hessian: H(u) ∈ (E1 −→ E1∗ ) is called the Hessian of J at u.
13.2.7 Mean value theorem
Let J be a functional as in 13.1. Let us assume that [u + tφ, t ∈ [0, 1]],
4
is contained in U . Let the function g : [0, 1] → be defined as
t −→ g(t) = J(u + tφ)
Let J (u + tφ, φ) exist. Then
J(u + (θ + t)φ) − J(u + tφ) g(t + θ) − g(t)
lim = lim = g (t)
θ→0 θ θ→0 θ
showing g is once differentiable in [0,1]
Thus, g (t) = J (u + tφ, φ) (13.8)
Similarly, if J (u + tφ, φ, φ) exist and J is twice differentiable then
g (t) = J (u + tφ; φ, φ) (13.9)
414 A First Course in Functional Analysis
13.2.8 Lemma
Let J be as in 13.1. Let u ∈ U , φ ∈ E1 be given. If [u+tφ : t ∈ [0, 1]] ∈ U
and J is once G-differentiable on this set in the direction, φ then there exists
a t0 ∈]0, 1[, such that
J(u + φ) = J(u) + J (u + t0 φ, φ) (13.10)
Proof: g : [0, 1] −→ 4 and g is differentiable in ]a, b[. The classical mean
value theorem yields:
g(1) = g(0) + 1 · g (t0 ), t0 ∈]0, 1[ (13.11)
Using the definition of g(t) and (13.8), (13.10) follows from (13.11).
13.2.9 Lemma
Let U and J be as in lemma 13.2.8. If J is twice G-differentiable on the
set [u + tφ; t ∈ [0, 1]] in the directions φ, ψ, then there exists a t0 ∈]0, 1[
such that
1
J(u + φ) = J(u) + J (u, φ) + J (u + t0 φ, φ, φ) (13.12)
2
Proof: The classical Taylor’s theorem applied to g on ]0, 1[, yields
1 2
g(1) = g(0) + 1 · g (0) + · 1 g (t0 ) (13.13)
2!
Using the definition of g(t), (13.8) and (13.9), (13.12) follows from
(13.13).
13.2.10 Lemma
Let E1 and E2 be two normed linear spaces. U an open subset of
E1 and let φ ∈ E1 be given. If the set [u + tφ; t ∈ [0, 1]] ∈ U and
P : U ⊂ E1 −→ E2 is a mapping which is G-differentiable everywhere on
the set [u + tφ; t ∈ [0, 1]] in the direction φ then, for any h ∈ E1∗ , there
exists a th ∈]0, 1[, such that
h(P (u + φ)) = h(P u) + h(P (u + th φ, φ) (13.14)
Proof: Let g : [0, 1] −→ 4 be set as
t −→ g(t) = h(P (u + tφ)) (13.15)
where h : U ⊂ E1 −→ . 4
Then g (t) exists in ]0, 1[ and
g(t + t ) − g(t) h(P (u + (t + t )φ) − h(P (u + tφ))
g (t) = Lt = Lt
t →0 t t →0 t
= h(P (u + tφ, φ) for t ∈]0, 1[,
since h is a linear functional defined on E1 .
Now, (13.14) follows immediately on applying the classical mean value
theorem to the function g.
Variational Problems 415
13.2.11 Theorem
Let E1 , E2 , u, φ and U be as in above. If P : U ∈ E1 −→ E2 is
G-differentiable in the set [u + tφ; t ∈ [0, 1]] in the direction φ, then there
exists a t0 ∈]0, 1[, such that,
with u = v.
416 A First Course in Functional Analysis
13.2.16 Theorem
4
If a functional J : E −→ is convex and admits of a gradient G(u) ∈ E ∗
at every u ∈ E, then J is weakly lower semicontinuous.
Proof: Let vn be a sequence in E, such vn − u is in E. Then
G(u)(vn − u) −→ 0 as n → ∞, since vn is weakly convergent in u. On
the other hand, since J is convex, we have by theorem 13.2.14,
13.2.17 Theorem
If a functional J : U ⊂ E −→ 4 on the open convex set of a normed
linear space, E is twice G-differentiable everywhere in U in all directions,
and if the form (φ, ψ) −→ J (u, φ, ψ) is non-negative, i.e., if J (u, φ, φ) ≥ 0
for all u ∈ U and φ ∈ E with φ = 0, then J is convex.
If the form (φ, ψ) −→ J (u, φ, ψ) is positive, i.e., if J (u, φ, φ) > 0 for
all u ∈ U and φ ∈ E with φ = 0, then J is strictly convex.
Proof: Since U is convex, the set [u + λ(v − u), λ ∈ [0, 1]] is contained in
U whenever u, v ∈ U . Then, by Taylor’s theorem [see Cea [11]], we have,
with φ = v − u,
1
J(v) = J(u) + J (u, v − u) + J (u + λ0 (v − u), v − u, v − u) (13.25)
2
1
J(v) = a(v, v) − L(v) for φ ∈ H, φ = θ.
2
1
J(v + φ) − J(v) = a(v + φ, v + φ) − L(v + φ)
2
1
− a(v, v) + L(v)
2
1 1 1
= a(v, φ) + a(φ, v) + a(φ, φ) − L(φ)
2 2 2
1
= a(v, φ) − L(φ) + a(φ, φ).
2
Variational Problems 419
13.3.4 Remark
The converse is not always true.
Example one in 13.2.3 has a G-derivative at (0, 0), but is not F-
differentiable.
13.4.2 Definition
A functional J on U is said to have a global minimum in U if there exist
a u ∈ U , such that
J(u) ≤ J(v) for all v ∈ U (13.29)
13.4.3 Theorem
Suppose E, U and J : U −→ 4 fulfil the following conditions:
1. E is a reflexive Banach space
2. U is weakly closed
3. U is weakly bounded and
4. J : U ⊂ E −→ 4 is weakly lower semicontinuous.
Then J has a global minimum in U .
Proof: Let m denote inf J(v). If vn is a minimizing sequence for J, i.e.,
v∈U
for all v ∈ E.
13.4.4 Theorem
If E, U and J satisfy the conditions (1), (2), (4) and J satisfy the
condition (5)
lim J(v) = +∞,
v→∞
1
J(v) = J(u) + J (u, v − u) + J (u + λ0 (v − u), (v − u), (v − w)), 0 < λ < 1.
2
1
J(v) = J(z) + G(z)(v − z) + J (z + λ0 (v − z), (v − z), (v − z))
2
for some λ0 ∈]0, 1[ (13.31)
Now, |G(z)(v − z)| ≤ ||G(z)|| ||v − z||, (13.32)
Condition (3) yields,
J (z + λ0 (v − z), (v − z), (v − z)) ≥ ||v − z||E(||v − z||) (13.33)
Using (13.32) and (13.33), (13.31) reduces to
1
J(v) ≥ J(z) + ||v − z|| E(||v − z||) − ||G(z)|| .
2
Here, since z ∈ U is fixed, as ||v|| −→ +∞
||v − z|| −→ +∞,
13.4.6 Theorem
Suppose U is a convex subset of a Banach space and J : U ⊂ E −→ 4
is a G-differentiable (in all directions) convex functional.
Then, u ∈ U is a minimum for J, (i.e., J(u) ≤ J(v) for all v ∈ E) if,
and only if u ∈ U and J (u, v − u) ≥ 0 for all v ∈ U .
Proof: Let u ∈ U be a minimum for J. Then, since U is convex,
u, v ∈ U =⇒ u +
n (v − u) ∈ U as
n → 0 for each n. Hence
J(u) ≤ J(u +
n (v − u))
J(u +
n (v − u)) − J(u)
Therefore, lim ≥ 0,
n →0+
n
i.e., J (u, v − u) ≥ 0 for any v ∈ E.
Conversely, since J is convex and G-differentiable by condition (i) of
theorem 13.2.14, we have
If we choose E(t) = αt, then all the assumptions of theorem 13.4.5 are
fulfilled by E, J and K so that problem (PI) has a unique solution.
Also, by theorem 13.4.6 the problem (PI) is equivalent to
(PII): To find
13.4.8 Theorem
(1) There exists a unique solution u ∈ K of the problem (PI).
(2) Problem (PI) is equivalent to problem (PII). The problem (PII)
is called a variational inequality associated to the closed, convex set and
the bilinear form a(·, ·).
The theorem 13.4.7 was generalized by G-stampaccia (Cea [11]) to the
non-symmetric case. This generalizes and uses the classical Lax-Miligram
theorem [see Reddy [45]]. We state without proof the theorem due to
Stampacchia.
13.4.9 Theorem (Stampacchia)
Let K be a closed convex subset of a Hilbert space H and a(·, ·) be a
bilinear bi-continuous coercive form (sec 11.7.2) on H. Then, for any given
L ∈ H, the variational inequality (13.34) has a unique solution u ∈ K.
For proof see Cea [11].
13.5 Distributions
13.5.1 Definition: Support
4
The support of a function f (x), x ∈ Ω ⊂ n is defined as the closure
4
of the set of points in n at which f is non-zero.
13.5.2 Definition: smooth function
4 4
A function φ : n −→ is said to be smooth or infinitely differentiable
if its derivatives of all order exist and are continuous.
13.5.3 Definition: C0∞ (Ω)
The set of all smooth functions with compact support in Ω ⊂ 4n is
denoted by C0∞ (Ω).
13.5.4 Definition: test function
A test function φ is a smooth function with compact support,
φ ∈ C0∞ (Ω).
13.5.5 Definition: generalized derivative
A function u ∈ C ∞ (Ω) is said to have the αth generalized derivative
α
D u, 1 ≤ |α| ≤ m, if the following relation (generalized Green’s
formula) holds:
Dα uφdx = (−1)|α| uD α φdx for every φ ∈ C0α (Ω) (13.35)
Ω Ω
where u and v along with their derivatives upto m, are square integrable
in the Lebesgue sense [see chapter 10].
|Dα u|2 dm < ∞ for all |α| ≤ p.
Ω
The space C ∞ (Ω) can be completed by adding the limit points of all
Cauchy sequences in C ∞ (Ω). It turns out that the distributions are those
limits points.
We can thus introduce the Sobolev space H 1 (Ω) as follows:
,
1 ∂v 2
H (Ω) = v : v ∈ L2 (Ω), ∈ L (Ω), j = 1, 2, . . . , p , (13.38)
∂xj
∂v
where Dj v = are taken in the sense of distributions,
∂xj
i.e., Dj vφdx = − vDj φdx for all φ ∈ D(Ω) (13.39)
Ω Ω
426 A First Course in Functional Analysis
where D(Ω) denotes the space of all C ∞ -functions with compact support
on Ω. H 1 (Ω) is provided with the inner product
n
u, v = u, vL2 (Ω) + Dj u, Dj vL2 (Ω) (13.40)
j=1
⎧ ⎫
⎨
n ⎬
= uv + (Dj u)(Dj v) dx (13.41)
Ω ⎩ ⎭
j=1
∂ n
= nj (x)D j (13.43)
∂n j=1
n
∂u
i.e., D j u, Dj v = (Δu)vdx + · vdσ (13.45)
j=1 Ω Γ ∂n
13.6.5 Remark
(i) u ∈ H 2 (Ω) =⇒ Δu ∈ L2 (Ω)
(ii) Since D j u ∈ H 1 (Ω) by trace theorem (13.6.3) γ(Dj u) exists and
belongs to L2 (Ω) so that
∂u
n
= nj γ(Dj u) ∈ L2 (Γ).
∂n j=1
Hence, by using the density and trace theorems (13.6.2 and 13.6.3),
the formula (13.4.3) is valid.
E = {v : v ∈ H 1 (Ω); γv = 0 on Γ1 } (13.46)
|L(v)| ≤ ||f ||L2 (Ω) ||v||L2 (Ω) ≤ ||f ||L2 (Ω) ||v||H 1 (Ω) for v ∈ E.
Then the problems (PI) and (PII) respectively become
(PIII) To find u ∈ E, J(u) ≤ J(v) for all v ∈ E (13.49)
(PIV) To find u ∈ E, u, φ = f, φL2 (Ω) for all v ∈ E (13.50)
Theorem 13.4.8 asserts that these two equivalent problems have unique
solutions.
The problem (PIV) is the Weak (or Variational) formulation of the
[34]]. Thus, if the problem (IV) has a regular solution then it is the solution
of the classical problem
⎧ ⎫
⎪
⎪ −Δu + u − f = 0 on Ω ⎪ ⎪
⎪
⎨ ⎪
⎬
u=0 on Γ1 (13.53)
⎪
⎪ ⎪
⎪
⎪
⎩
∂u
=0 on Γ2 ⎭ ⎪
∂n
13.6.8 Remark
The variational formulation (PIV) is very much used in the Finite
Elements Method.
CHAPTER 14
THE WAVELET
ANALYSIS
The concept of Wavelet was first introduced around 1980. It came out
as a synthesis of ideas borrowed from disciplines including mathematics
(Calderón Zygmund operators and Littlewood-Paley theory), physics
(coherent states formalism in quantum mechanism and renormalizing
group) and engineering (quadratic mirror filters, sidebend coding in signal
processing and pyramidal algorithms in image processing) (Debnath [17]).
Wavelet analysis provides a systematic new way to represent and analyze
multiscale structures. The special feature of Wavelet analysis is to
generalize and expand the representations of functions by orthogonal
basis to infinite domains. For this purpose, compactly supported
[see 13.5] basis functions are used and this linear combination represents
the function. These are the kinds of functions that are realized by physical
devices.
There are many areas in which wavelets play an important role, for
example
430
The Wavelet Analysis 431
14.2.4 Orthogonality
The terms in a wavelet series are orthogonal to one another, just like
the terms in a Fourier series. This means that the information carried by
one term is independent of the information carried by any other.
14.2.5 Multiresolution representation
Multiresolution representation describes what is called a hierarchical
structure. Hierarchical structures classify information into several
categories called levels or scales so that, higher in the hierarchy a level
is, the fewer the number of members it has. This hierarchy is prevelant
in the social and political organization of the country. Biological sensory
system, such as visions, also have this hierarchy built in. The human
vision system provides wide aperture detection (so events can be detected
early) and high-resolution detection (so that the detailed structure of the
visual event can be seen). Thus, a multiresolution or scalable mathematical
representation provides a simpler or more efficient representation than the
usual mathematical representation.
14.2.6 Functions and their representations
Representation of continuous functions
Suppose we consider a function of a real variable, namely,
f (x) = cos x, x ∈ 4
where 4 denotes the continuum of real numbers. For each x ∈ , we 4
have a definite value for cos x. Since cos x is periodic, all of its values are
determined by its values on [0, 2π[. The question arises as to how best we
can represent the value of the function cos x at any point x ∈ [0, 2π[. There
is an uncountable number of points in [0, 2π[ and an uncountable number
of values of cos x, as x varies. If we represent
∞
x2n
cos x = (−1)n (14.1)
n=0
2n!
+
Supposethat {cn } is a given discrete sequence of complex numbers in l2 ( ),
that is |cn |2 < ∞, then we define the Fourier transform of the sequence
f = {cn } to be the Fourier series [see 14.3],
fˆ(ξ) = cn e2πinξ
n
where ⎛ ⎞
a0p a0p+1 ··· a0p+m−1
⎜ .. ⎟
Ap = ⎜
⎝ .
⎟
⎠
am−1
p am−1
p+1 ··· m−1
ap+m−1
where we assume that An1 and An2 are both non-zero matrices.
g = n2 − n 1 + 1 (14.13)
i.e., the number of terms in the series (14.12) is called the genus of the
Laurent series A(z) and the matrix A.
14.3.2 Definition: adjoint A(z) of the Laurent series A(z)
Let,
A(z) = A∗ (z −1 )
= A∗p z −p (14.14)
p
14.3.7 Examples
1. Haar matrix of rank 2
1 1
Let A1 = . Here A1 AT1 = 2I.
1 −1
Sum of the elements of the first row = 2 = rank of A.
Sum of the elements of the second row = 0.
Hence, A1 is a wavelet matrix of rank 2.
1 1
Similarly, A2 = , A2 AT2 = 2I and fulfils conditions (14.15)
−1 1
and (14.16).
Hence, A2 is a wavelet matrix of rank 2 too.
The general complex Haar wavelet matrix of rank 2 has the form
1 1
, θ ∈ 4.
−eiθ eiθ
1 1+ 3 3+ 3 3− 3 1− 3
D2 = √ √ √ √ (14.18)
4 −1 + 3 3 − 3 −3 − 3 1 + 3
The Wavelet Analysis 437
D2 D2T = 2I.
m−1
m−1
Proof: From (14.16) and (14.15), we have hj hj = m and hj = m.
j=0 j=0
m−1
It follows that hj = m.
j=0
m−1
m−1
Now, |hj − 1|2 = (hj hj − hj − hj + 1)
j=0 j=0
m−1
m−1
m−1
m−1
= hj hj − hj − hj + 1
j=0 j=0 j=0 j=0
=m−m−m+m
=0
α+β+γ =0
α2 + β 2 + γ 2 = m.
If we take β = γ then α + 2β = 0
α2 + 2β 2 = 6β 2 = m
440 A First Course in Functional Analysis
; ;
m m
i.e., β = , α = −2
6 6
Hence, the last but one row is
; ; ;
m m m
0, 0, . . . , 0, −2 , , .
6 6 6
Similarly, let the rotation yield for the (m − s)-th row all the first
(m − s − 1) elements as zeroes and the last (s + 1) elements as non-zeroes.
Let these be α1 , α2 , . . . , αs , αs+1 .
Then αi = 0, αi2 = m.
i i
Taking α2 = α3 = · · · = αs = αs+1
we have α1 = −sα2
;
m
αi2 = m or, s2 α22 + sα22 =m or, α2 = .
s2 + s
i
Hence, the (m − s)-th row is
; ; ;
m m m
0, 0, 0, . . . , −s , ,..., .
s2 + s s2 + s s2 + s
Thus we get the expression for 0 [see 14.21].
14.4.3 The wavelet function
If φ is a scaling function for the wavelet matrix A, then the wavelet
functions {ψ 1 , ψ 2 , . . . , ψ m−1 } associated with matrix A and the scaling
function φ are defined by the formula
mg−1
ψ s (x) = asj φ(mx − j) (14.25)
j=0
14.4.4 Theorem
+
Let A ∈ W M (m, g; ) be a wavelet matrix. Then, there exists a unique
4
φ ∈ L2 ( ), such that
14.4.5 Examples
1. The Haar functions If 0= 1 1
−1 1
is the canonical Haar wavelet
matrix of rank 2, then the scaling function φ satisfies the equation
φ(x) = φ(2x) + φ(2x − 1) (14.26)
Hence, φ(x) = χ[0, 1[ where χK , the characteristic function of a subset K
[see 10.2.10] is a solution of the equation (14.26) [see 14.2(b)].
The wavelet function ψ = −φ(2x) + φ(2x − 1), where φ(x) = χ[0, 1[.
For graph, see figure 14.2(c).
1.5
0.5
−0.5
−1.5
−0.5 0 0.5 1 1.5
0.5
−0.5
−1.5
−0.5 0 0.5 1 1.5
1 1+ 3 3+ 3 3− 3 1− 3
Let D2 = √ √ √ √
4 −1 + 3 3 − 3 −3 − 3 1 + 3
This is a wavelet matrix of rank 2 and genus 2 and discovered by
Deubechies. For graphs of the corresponding scaling and wavelet functions,
see Resnikoff and Wells [46]. The common support of φ and ψ is [0.3].
We end by stating a theorem due to Lawton, (Lawton [46]).
14.4.6 Theorem
+
Let A ∈ W M (m, g; ). Let W (A) be the wavelet system associated
4
with A and let f ∈ L2 ( ) (3.1.3). Then, there exists an L2 -convergent
expansion
∞
m−1 ∞ ∞
f (x) = cj φj (x) + dsij ψij
s
(x) (14.27)
j=−∞ s=1 i=0 j=−∞
DYNAMICAL SYSTEMS
x
O π 2π
Fig. 15.1
443
444 A First Course in Functional Analysis
Given a dynamical system on X, the space X and the map are respectively
called the phase space and the phase map (of the dynamical system).
Unless otherwise stated, X is assumed to be given.
Dynamical Systems 445
4
for all x1 , x2 ∈ n and a given positive number M , so that the conditions
on solutions of (15.4) are obtained. We next want to show that the map
4 4 4
π : n × → n such that π(x, t) = ψ(t, x) defines a dynamical system
4
on n .
For that, we note that
π(x, 0) = ψ(0, x) = x
π(π(x, t1 ), t2 ) = ψ(t2 , ψ(t1 , x)) = ψ(t1 + t2 , x)
= π(x, t1 + t2 )
a maximal interval (a(x), b(x)), −∞ a(x) < 0 < b(x) +∞. For each
x ∈ D, define Γ+ (x) = {ψ(t, x) : 0 t < b(x)} and Γ− (x) = {ψ(t, x) :
a(x) t 0}, Γ+ (x) and Γ− (x) are respectively called the positive and
the negative trajectories respectively through the point x ∈ D.
We will show that to each system (15.6), there corresponds a system
dx
dt
= ẋ = F (x), x ∈ 4n (15.7)
4
where F : D → n such that (15.7) defines a dynamical system on D with
the property that for each x ∈ D the systems (15.6) and (15.7) have the
same positive and the same negative trajectories.
4
If D = n , then given the equation (15.4), we set,
dx F (x)
= ẋ(t) = F1 (x) = (15.8)
dt 1 + ||F (x)||
where || · || is the Euclidean norm. If D = 4n , then the boundary ∂D of
D = Φ and is closed.
We next consider the system
dx F (x) ρ(x, ∂D)
= ẋ(t) = F1 (x) = · (15.9)
dt 1 + ||F (x)|| 1 + ρ(x, ∂D)
where ρ(x, ∂D) = inf{||x − y|| : y ∈ ∂D}.
In other words, ρ(x, ∂D) is the distance of x from ∂D.
Since f satisfies Lipschitz condition, equation (15.4) has a unique
solution.
0 0
0 F (x) F (y) 0
Now, ||F1 (x) − F1 (y)|| = 0 0 − 0
1 + ||F (x)|| 1 + ||F (y)|| 0
1
≤ ||F (x) − F (y)|| + ||F (x)|| ||F (y)||×
(1 + ||F (x)||)(1 + ||F (y)||)
0 0
0 F (x) F (y) 0
0 0
0 ||F (x)|| − ||F (y)|| 0
yi = Fi (x1 , x2 , . . . , xn ), i = 1, 2, . . . , m (15.10)
15.2.6 Remark
Note that G : U → V is a diffeomorphism if and only if, m = n and the
matrix of partial derivatives,
∂Gi
G (x1 , . . . , xn ) = , i, j = 1, . . . , n
∂xj
is non-singular at every x ∈ U .
T
Example: Let G(x, y) =
exp y
exp x
4
with U = 2 and V = {(x, y) : x >
0, y > 0}.
0 exp y
G (x, y) = , det G (x, y) = − exp(x + y) = 0 for each
exp x 0
4
(x, y) ∈ 2 .
Thus, G is a diffeomorphism.
15.2.7 Two types of dynamical systems
We note that in a Dynamical system the state changes with time {t}.
The two types of dynamical system encountered in practices are as follows:
(i) xt+1 = G(xt ), t ∈ Z or N (15.11)
Such a system is called a discrete system.
(ii) When t is continuous, the dynamics are usually described by a
differential equation,
dx
= ẋ = F (x) (15.12)
dt
In (15.11), x represents the state of the system and takes values in the
state space or phase space X [see 15.1.1]. Sometimes the phase space is the
Euclidean space or a subspace of it. But it can also be a non-Euclidean
structure such as a circle, sphere, a torus or some other differential
manifold.
15.2.8 Advantages of taking the phase space as a differential
manifold
If the phase space X is Euclidean, then it is easy to analyse. But if the
phase space X is non-Euclidean but a differential manifold there is also an
advantage. This is because a differential manifold is ‘locally Euclidean’ and
this allows us to extend the idea of differentiability to functions defined on
them. If Y is a manifold of dimension n, then for any x ∈ Y we can find a
neighbourhood Nx ⊆ Y containing x and a homomorphism h : Nx → Rn
which maps Nx onto a neighbourhood of h(x) ∈ Rn . Since we can define
coordinates in U = h(Nx ) ⊆ Rn (the coordinate curves of which can be
mapped back onto Nx ) we can think of h as defining local coordinates on
the patch Nx of Y [see figure 15.2].
Dynamical Systems 449
The pair (U, h) is called a chart and we can use it to give differentiability
on Nx . Let us assume that G : Nx → Nx then G induces a mapping
Ĝ = h · G · h−1 : U → U [see figure 15.3]. We say G is a C k -map on Nx if
Ĝ is a C k -map on U .
z
θ
Fig. 15.2 Cylinder
Nx G Nx
h−1 h
U ∧
U
G
Fig. 15.3
15.3.4 Remark
1. A fixed point is a periodic point of period one.
2. Since (Gp )q (x∗ ) = Gp (x∗ ), q ∈ Z for a periodic point x∗ of G with
period p, x∗ is a fixed point of Gp (x∗ ).
3. If x∗ is a periodic point of period p for G, then all other points in the
orbit of x∗ are periodic points of period p of G. For if Gp (x∗ ) = x∗ ,
then Gp (Gi (x∗ )) = Gi (Gp (x∗ )) = Gi (x∗ ), i = 0, 1, 2, . . . , q − 1.
Gm (x∗ ) ⊆ Nx∗ for all m > 0. The above definition implies that iterates
of points near to a stable fixed point remain near to it for m ∈ Z+ . This
is in conformity with the definition of a stable equilibrium point for a
moving particle.
15.3.6 Remark
1. If a fixed point x∗ is stable and limm→∞ Gm (x∗ ) = x∗ , for all x
in some neighbourhood of x∗ , then the fixed point is said to be
asymptotically stable.
2. Unstable point. A fixed point x∗ is said to be unstable if for every
neighbourhood Nx∗ of x∗ Gm (x) ∈ Nx∗ for all x ∈ Nx∗ .
15.3.7 Example
We refer to the example in 15.3.2. To determine stability, we plot x2 − 1
and then sketch the vector field [see figure 15.4]. The flow is to the right
where x2 − 1 > 0 and to the left where x2 − 1 < 0. Thus, x∗ = −1 is stable
and x∗ = 1 is unstable.
Dynamical Systems 451
•
x
x
O
Fig. 15.4
15.4.2 Remark
1. These theorems have deep implications. Different trajectories do
not intersect.
For if two trajectories intersect at some point in the phase space, then
starting from the crossing point we get two solutions along the two
trajectories. This contradicts the uniqueness of the solution.
2. In two dimensional phase spaces let us consider a closed orbit C in the
phase plane. Then any trajectory starting inside C will always lie within
C. If there are fixed points inside C, then the trajectory may approach one
of them.
But if there are no fixed points inside C, then by intuition we can
say that the trajectory can not move inside the orbit endlessly. This is
supported by the following famous theorem.
Poincaré-Bendixon theorem:
If a trajectory is confined to a closed, bounded region and there are no
fixed points in the region, then the trajectory must eventually approach a
closed circuit (Arrowsmith and Pace [3]).
452 A First Course in Functional Analysis
Fig. 15.5
4
For μ < 0, the equation (15.14) yields ẋ < 0 for all x ∈ . When μ = 0
there is a non-hyperbolic fixed point at x = 0, but ẋ < 0 for all x = 0
(Arrowsmith and Place [3]). For μ > 0, μ − x2 = 0 ⇒ x = ±μ1/2 for
x > μ1/2 , ẋ < 0 and for x < μ1/2 , ẋ > 0. Hence, x = μ1/2 is a stable fixed
point. On the other hand, x = −μ1/2 is an unstable fixed point.
Example 2. F (μ, x) = μx − x2 (15.15)
If μ > 0, x = μ is stable and x = 0 is unstable. The stabilities are
reversed when μ < 0. At μ = 0, there is one singularity at x = 0 and ẋ < 0
Dynamical Systems 453
for all x = 0. This leads to a bifurcation as depicted below [see figure 15.6].
Fig. 15.6
List of Symbols
454
List of Symbols 455
Inf Infimum
Sup Supremum
· Norm
456 A First Course in Functional Analysis
xm → x {xm } tends to x
lim sup Limit supremum
lim inf Limit infimum
· 1 l1 —norm
· 2 l2 —norm
· ∞ l∞ —norm
∞
xn Summation of xn for n = 1, . . . ∞
n=1
m
xn Summation of xn for n = 1, 2, . . . m (finite)
n=1
(n)
lp n-dimensional pth summable space
X0 The interior of the set X
c0 Space of all sequences converging to 0
x + Lq Quotient norm of x + L
Span L Set of linear combinations of elements in L
σa (A) Approximate eigenspectrum of an operator A
rσ (A) Spectral radius of an operator A
T
A Transpose of a matrix A
D(A) Domain of an operator A
R(A) Range of an operator A
N (A) Null space of an operator A
BV ([a, b]) Space of scalar-valued functions of bounded
variation on [a, b]
N BV ([a, b]) Space of scalar-valued normalized functions of
bounded variation on [a, b]
Δ Forward difference operator
w
xn −→ x {xn } is weak convergent to x
x, y Inner product of x and y
x⊥y x is orthogonal to y
E⊥F E is orthogonal to F
T
M Conjugate transpose of a matrix M
A≥0 A is a non-negative operator
w(A) Numerical range of an operator A
List of Symbols 457
m(E)
Lebesgue measure of a subset E of 4
x dm Lebesgue integral of a function x over a set E
E
Var (x) Total variation of a scalar valued function x
essupE |x| Essential supremum of a function |x| over a
set E
L∞ (E) The set of all equivalent classes of essentially
bounded functions on E
Span L Closure of the span of L
⊥
M Set orthogonal to M
⊥⊥
M Set orthogonal to M ⊥⊥
(Ex → Ey ) Space of all bounded linear operators mapping
n.l.s. Ex → n.l.s. Ey
H Hilbert space
−1
A Inverse of an operator A
∗
A Adjoint of an operator A
A Closure of A
Aλ Operator of the form A − λI
A Norm of A
A(x, x) Quadratic Hermitian form
A(x, y) Bilinear Hermitian form
k
C ([0, 1]) Space of continuous functions x(t) on [0, 1] and
having derivatives to within k-th order
Ex∗ Conjugate (dual) of Ex
Ex∗∗ Conjugate (dual) of Ex∗
ΠEx (x) Canonical embedding of Ex into Ex∗∗
χE Characteristic function of a set E
δij Kronecker delta
sgn z Signum of z ∈ +
{e1 , e2 , . . .} Standard Schauder basis for 4n or +n or lp , 1 ≤
p<∞
Gr(F ) Graph of a map F
I Identity operator
ρ(A) Resolvent set of an operator A
σ(A) Spectrum of an operator A
458 A First Course in Functional Analysis
Abbreviations
BVP Boundary value problems
LHS Left hand side
ODE Ordinary differential equations
RHS Right hand side
s.t. Such that
WRT With regards to
a.e. Almost everywhere
Bibliography
459
460 A First Course in Functional Analysis
463
464 A First Course in Functional Analysis