(Graduate Studies in Mathematics, V. 50) Y. A. Abramovich, Charalambos D. Aliprantis-An Invitation To Operator Theory-Amer Mathematical Society (2002)
(Graduate Studies in Mathematics, V. 50) Y. A. Abramovich, Charalambos D. Aliprantis-An Invitation To Operator Theory-Amer Mathematical Society (2002)
(Graduate Studies in Mathematics, V. 50) Y. A. Abramovich, Charalambos D. Aliprantis-An Invitation To Operator Theory-Amer Mathematical Society (2002)
Operator Theory
Y. A. Abramovich
C. D.Aliprantis
Graduate Studies
in Mathematics
Volume SO
Y. A. Abramovich
Indiana University-Purdue University Indiana po/ls
C. D. Aliprantis
Purdue University
Graduate Studies
in Mathematics
Volume 50
2000 Mathematics Subject Classification. Primary 46Axx, 46Bxx, 46Gxx, 47Axx, 47Bxx,
47Cxx, 47Dxx, 47Lxx, 28Axx, 28Exx, 15A48, 15A18.
QA329.A25 2002
2002074420
Copying and reprinting. Individual readers of this publication, and nonprofit libraries
acting for them, are permitted to make fair use of the material, such as to copy a chapter for use
in teaching or research. Permission is granted to quote brief passages from this publication in
reviews, provided the customary acknowledgment of the source is given.
Republication, systematic copying, or multiple reproduction of any material in this publication
is permitted only under license from the American Mathematical Society. Requests for such
permission should be addressed to the Acquisitions Department, American Mathematical Society,
201 Charles Street, Providence, Rhode Island USA. Requests can also be made by
e-mail to reprint-permission@ams. org.
2002 by the American Mathematical Society. All rights reserved.
The American Mathematical Society retains all rights
except those granted to the United States Government.
Printed in the United States of America.
The paper used in this book is acid-free and falls within the guidelines
established to ensure permanence and durability.
Visit the AMS home page at http://www.anis.org/
10987654321 070605040302
To Alla, Julia, and Jane
YAA
Foreword ix
Chapter 1. Odds and Ends 1
V
- Contents
Bibliography 505
Index 521
Foreword
ix
- - -
Foreword
From the early 1960's it was realized that the topological and order
structures of operators were related and that they should be studied to-
gether. Since then much research has aimed at producing a theory that
would encompass these structures in a unified manner. The "unification"
of the order and topological properties of operators is due to the efforts of
many mathematicians in various countries. The period of this unification
can be roughly divided into two parts; the pre-1980 period and the post4 980
period. Let us describe here the basic achievements done in each period.
The pre-1980 period can be described as the period where most of the
foundations of the order structure of vectors spaces and the lattice proper-
ties of operators were laid down. In this era the theory of partially ordered
vector spaces was developed through the efforts of mathematicians from
many countries. There were several monographs written on the subject.
Those written by A. L. Peressini [264] and G. Jameson [161] and the last
part of H. H. Schaefer's book [289] were devoted entirely to ordered vec-
tor spaces. The books by W. A. J. Luxemburg and A. C. Zaanen [2253,
B. Z. Vulikh [328], and S. Kaplan [176] studied the order structure of Riesz
spaces, while D. H. Fremlin [128] and C. D. Aliprantis and 0.
shaw [29] studied topological properties of Riesz spaces. Monographs on
Banach lattices and positive operators were written by M. A. Krasnoselsky
et al. {191], H. J. Krieger [194], H. E. Lacey [197], J. Lindenstrauss and
L. Tzafriri [205, 206], H. H. Schaefer [290], and H.-U. Schwarz [298].
The post-1980 period is characterized by the study of positive operators
from all points of view. Special attention was paid to the properties of op-
erators dominated by positive operators having some compactness property.
We also find in this era a complete study of integral operators in terms of
their lattice structure. Monographs regarding positive operators were writ-
ten in this period by C. D. Aliprantis and 0. Burkinshaw [30], W. Arendt et
al. [41], A. V. Bukhvalov et al. [78], P. Meyer-Nieberg [240], and A. C. Za-
anen [343, 344]. The list of very recent monographs on the subject of
positive operators includes the books by Y. A. Abrarnovich, E. L.
son, and A. K. Kitover [21], Y. A. Abramovich and A. K. Kitover [22],
A. G. Kusraev [195], and W. Wnuk [338].
Besides its important internal role in mathematics, the theory of Riesz
spaces and positive operators has important applications to several disci-
plines. The areas where the theory of Riesz spaces and positive operators
has been found to be useful include: game theory [182]; finance [151, Chap-
ter 31]; economics [26]; nuclear reactor theory [64, 65]; statistical decision
theory [201]; and structured population dynamics [80, 248].
The objective of this book is to present at the graduate level in a self-
contained manner the theory of linear operators between Banach spaces
Foreword
indebted to W. Geller who read most of the book and, apart from fixing
some mathematical irregularities, corrected a great number of irregularities
in our English. Without his help our writing would have been considerably
less idiomatic.
Finally, we would like to thank our wives, Alla and Bernadette, for
coping with our idiosyncrasies during the many years it took us to assemble
the material, write, and typeset this text. To accomplish this project without
their patience and understanding would not have been possible.
'The symbols R and C will be used exclusively to denote the set of real and complex numbers,
respectively. Also, as usual, N will denote the set of natural numbers, Q the set of rational numbers,
and Z the set of integers.
1.1. Banach Spaces, Operators, and Linear Functionals 3
(T**x**,y*) = (x**,T*y*)
for all E X** and all E is called the double adjoint of the
operator T. it is not difficult to see that [ITCI = IIT* = Indeed,
= sup [
sup (y*Tx)]
IIxIl1
sup = 11TH.
IxLl1
From now on, unless otherwise stated, X and Y will denote two arbitrary
normed spaces. A subset A of (X, Y) is said to be pointwise bounded
if for each x E X there exists some > 0 (depending upon x) satisfying
for eachTEA.
Theorem 1.3 (The Uniform Boundedness Principle). Assume that X is a
Banach space and Y is a normed space. Then a subset of (X, Y) is norm
bounded if and only if it is pointviise bounded.
lizil = sup
OE[O,2ir]
IITCM = ITH.
Proof. First note that the inequality ITch IT]! is obvious. On the other
hand, for each Z = x + zy and any angle 9 we have
(Tx) cos 9 + (Ty) sin = IIT(x cos 9 + y sin 0)
IITII(IlxcosO+ ysin9jl)
<
This implies ITCZI IIZII or ITch So, ITch TI.
In the sequel, quite often, we shall identify T with without any men-
tion. When we do so the identification will be clear to the reader from the
context.
It is very common to denote the vectors x + of as column vectors
[x].
With this notation in mind, it is not difficult to see that every operator
T: can be represented by a unique operator matrix of the form
T T-S where 8, T: X X are linear operators, and as usual
= T
[x]
T(x + 2y)
= =
= Tx - Sy + 2(Sx + Ty).
Notice that in this notation, if T: X X is an operator, then the operator
is represented by the matrix
TO
Any operator T: for which S = 0 is called a real operator
on Of course, an operator T: is real if and only if T = for
some operator T: X X. Also, it should be noticed that for an operator
T: to be real it is necessary and sufficient that T leaves X
invariant, i.e., T(X) c x.
Thus, every operator T E can be identified with a matrix and
also with a pair of operators T, S E (X, Y). Equivalently, every operator T
1.1. Banach Spaces, Operators, and Linear Functionals - 7
Ref, i.e., g = Ref, and called the real part of f) such that
f(x) g(x) zg(zx) (*)
- = =
1. Odds and Ends
Exercises
1. If two vectors x and y in a normed space satisfy Ix + = lxii +
then show that = aiIxlI+1311y11 for all scalars a,/3 0. [HINT:
If the scalars a and /3 satisfy a /3, then note that
+ +
IIZlIc
e e' e
=
12. Recall that a (not necessarily surjective) operator T: X f Y between
normed spaces is said to be an isometry (or more precisely a linear
isometry) if IITXII = holds for each x E X.
Show that a bounded operator T: X * Y between two normed spaces
is an isometry if and only if its double adjoint T**: y** is likewise
an isometry. [HINT: Assume that T: X Y is an isometry. Clearly,
= IIT**M = 1 and so IIT**x**M for all E X**. Now fix
X* with 1. Then the formula y*(Tx) = defines a
continuous linear functional on the range of T such that 1. Let
y* also denote a continuous linear extension of to all of Y such that
<1; clearly T*y* = Now note that for each we have
= I(T*y*,x**)I =
IT**x**II
='
The mapping is called the nearest point mapping from X to A.
(d) Let C be a non-empty closed convex subset of a uniformly convex
Banach space X. Show that the nearest point mapping 7r is contin-
uous.
18. For a non-empty closed convex subset C of a Hilbert space X establish
the following properties of the nearest point mapping 7r: X + C.
(a) If s E C, then 71(5) = s.
(b) If s C, then 7r(s)E 5C.
(c) ForeachxEXandeachyECwehave
(d) For each x,y E X we have (7r(x) 'ir(y),x y) 0.
(e) For all s,y E X we have 1171(x) ir(v)1I S
(f) If C is a closed vector subspace of X, then for each s E X the vector
x 71(5) is orthogonal to C, i.e., (x 'ir(x), c) =0 for all c E C.
Remark: Recall that a non-empty subset A of a topological space is
i.e., if every vector can be written as a difference of two positive vectors (or
(1) x and j. x.
(2) For each fi E B and 'y E r there exists some ao E A (depending on
fi and 'y) satisfying z7 for all a ao.
The element x is called the order limit of the net
The following observation regarding order limits will be useful. If in the
definition of x we consider the product index set A = B x r and
define the two nets and by
and
o A net in a partially ordered set can have at most one order limit.
o If a net in a partially ordered set satisfies (resp.
and x (resp. x).
(See Exercise 4 at the end of the section.) The next result deals with order
convergence in Riesz spaces.
The above suprema and infima exist since the net is order bounded
and E is Dedekind complete. Clearly, and Moreover, since for
18 1. Odds and Ends
According to Lemma 1.20, for cr-Dedekind complete Riesz spaces this defi-
nition is, of course, equivalent to the general definition of order convergence
of nets as given in Definition 1.18.
Definition subset D of a partially ordered set is said to be order
1.22. A
closed (reap. cr-order closed) if D and xE D
(reap. D and Xn x imply x E D).
The principal ideals will play an important role in this book. A vector e> 0
is called an order unit (or a strong unit or simply a unit) if Ee = E, i.e.,
if for each x E X there exists some 0 such that <,\e.
An order closed ideal is called a band. Since a solid subset D in a Riesz
space E is order closed if and only if D fl E imply
x E D (see Exercise 5 at the end of the section), it follows that an ideal A
IximplyxeA.
If D is a subset of a Riesz space E, then the band generated
by D is the smallest (with respect to inclusion) band BD that contains D.
If D = {x}, then = is called the principal band generated by the
vector x. The band generated by an ideal is described as follows.
1.2. Banach Lattices and Positive Operators 21
Recall also that a Riesz space E is said to have the countable sup
property if for every subset D of E having a supremum there exists an at
most countable subset C of D satisfying sup C = sup D. The countable sup
property allows us to work with sequences rather than nets. For example,
if E satisfies the countable sup property, then F) = F).
A seminorm p on a Riesz space is said to be a lattice seminorm when-
ever implies p(x) p(y). A normed Riesz space is a Riesz space
equipped with a lattice norm. A normed Riesz space which is also norm
complete (i.e., a Banach space) is called a l3anach lattice. In a normed
Riesz space, the lattice operations are uniformly continuous functions. In
particular, this implies that every band is a closed subspace. Every band
projection on a normed Riesz space is always a contraction. We also have
the following simple but very useful result.
Lemma 1.30. In a normed Riesz space P2:
(a) The cone E+ is norm (and hence weakly) closed.
(b) The order intervals are norm (and hence weakly) closed.
(c) If a net E satisfies and 0, then x.
Our next result informs us that positive operators between Banach lat-
tices are continuous; for a proof see [30, p. 175].
Theorem 1.31. Every order bounded operator from a Banach lattice to a
normed Riesz space is continuous (and hence every positive and every regular
operator on a Banach lattice is also continuous)
Now consider two Banach lattices E and F with F Dedekind complete.
For a regular operator T: E F its r-norm is defined by ITlir = ITt 1
Note that we have
IiTIir = inf{II Sit: T S }.
This formula allows us to define the r-norm for regular operators with values
in a non-Dedekind complete Banach lattice. For the proof of the next result
see [30, Theorem 15.2, p. 248].
Theorem 1.32. If E and F are two Banach lattices with F Dedekind com-
plete, then the Dedekind complete Riesz space 4(E, F) of all regular oper-
ators from E to F equipped with the r-norm is a l3anach lattice.
= T(x)-f-T(y)--2T(xAy)
= T(x) + T(y) 2[T(x) A T(y)] = IT(x) T(y)I.
Letting y = 0 gives IT(x)1 = for all x E E.
24 1. Odds and Ends
T(xVy) = T(x)+T(y)T(xAy)
= T(x) + T(y) [T(x) A T(y)] = T(x) V T(y).
Exercises
1. Let C be a cone in a real vector space X. Show that the relation on
X defined by w y whenever s y E C is a vector space order such that
x+C.
2. Prove Theorem 1.15.
3. Show that order complete Riesz spaces and normed Riesz spaces are Ar-
chimedean.
4. Establish the following properties of order convergence.
(a) A net in a partially ordered set can have at most one order limit.
(b) If 5a s or Xa s in a partially ordered set, then Xa S.
(c) If a net {Sa}aEA in a partially ordered set satisfies Wa (resp. 5a .1.)
and 5a x, then 5a x (resp. j x).
= and Zn
and note that both sequences are increasing norm Cauchy sequences and
Xn+1 = Zn.]
17. Prove Theorem 1.32.
18. Give an example of a regular operator T: E + F between two Banach
lattices with F Dedekind complete satisfying 11Th <
19. Show that a band projection P on a normed Riesz space is a contraction
IPII 1).
20. If E is a normed Riesz space, then show that II = = I II
fx*(x)h x*h(ixi) 1
x*t 11 lIxil
we infer that IIx*11 1
On the other hand, if Ixjh I and e > 0
then from x*i(hxh) = sup{lx*(y)h: lxi we see that there
exists some y E E with lxi such that Ix*t(hxh) <
Therefore,
II
Ix*H1 =sup{Ix*1(hxt): lixil 1} tx*1I+E,
for each e > 0. This implies II
= j_i)
n+p1 n+p1
D={XEX: suchthat
then D is a countable set that is dense in X. This means that only sepa-
rable Banach spaces can possibly have bases. The basis problem was the
following long standing question: Does every separable Banach space have a
basis? In 1972 P. Enflo [120] produced a separable Banach space without
a basis, and so he provided a negative answer to the basis problem.
Now let be a basis in a Banach space X, and for each vector
x E X let x = be its series representation. The
functional of the basis is the linear functional defined by
= =
= =
IUxffl = sup
nEN i=i
(b) Since .
Ills equivalent to II there exists a constants C > 0 such
that IIxlII for each x E X. This implies
= HIXHI
= = =
Next, we shall prove that every Banach space has a basic sequence. To
establish this important result, we need a lemma due to S. Mazur.
Lemma L42 (Mazur). Let X be an infinite dimensional Banach space, let
Y be a finite dimensional vector subspace of X, and let 0. Then there
is a unit vector x e X such that (1 + + for all y E Y and
all scalars
Proof. Assume 0 < 1.It suffices to show that there exists some unit
vector x E X such that I (1 + ) y E Y with = 1.
Since the closed unit ball of Y is norm compact, there exist unit vectors
yi,. .
. , in Y such that for each unit vector y e Y there exists some i so
that < For each 1 i ri. choose some e X* satisfying
= I and X is infinite dimensional we can find
a unit vector x e X such that = 0 for each I i n., that is,
xe Ker Now if a is any scalar and y e Y satisfies = 1, then
choose some I <i <ri with YiM and note that
+ - + - =1-
Therefore, I (1 + + for all unit vectors y e Y, as desired. D
Proof1 Let X be an infinite dimensional Banach space. Fix some real num-
ber 0 and then choose a sequence of positive real numbers such
that K = + I + . Pick any unit vector E X and then
use Mazur's Lemma 1.42 to select inductively a sequence of unit vectors
X2, X3,.. such that for each ri. we have
IzII < (1 + +
for all scalars and all z in the vector subspace generated by {xi,... ,
Now note that if i-n > n., then for any choice of scalars cvi,.. . , we have
n m
K) =
= K) = m
m
m
m
lxii <
The proof of the next result that characterizes the unconditionally con-
vergent series is left to the reader.
Lemma 1.45. For a series in a Banach space the following state-
ments are equivalent.
(1) The series converges unconditionally.
(2) For any sequence of signs (i.e., 1 for each n) the series
Snxn is norm convergent.
(3) For every strictly increasing sequence of natural numbers the
series is norm convergent.
(4) For each E> 0 there exists a natural number k such that for each
finite subset A of N with mm A 2 k we have
xn a Banach space converges uncon-
ditionally, then xn = for each permutation ir of N.
=
where x= It should be clear thatis a linear operator. If
denotes the sequence of biorthogonal functionals with respect to the
basis and X X is the (bounded) finite-rank operator defined by
= then we have
=
Ap={mEN: and
and note that for each n n0 we have
k+p
= < 2E.
ik
M8(x)jj
______________________
iEF iEF
Proof. Assume that F and the scalars cEj and 'yj are as in the statement
of the lemma. Choose a continuous linear functional E X* of norm one
satisfying (>iEF = Next, pick any sequence of
signs s (81, 82,...) such that for each i e F we have = 1 if 0
and = 1 if <0. This implies
c
iEF iEF
as desired. D
Proof Define X =
[O,oo) by IlxMi lIMsxIl. It is easy to
verify that this formula defines a norm on X. Moreover, note that
lxiii
for each x E X. This shows that
are equivalent norms. To
and
finish the proof, notice that the unconditional constant of with respect
to the norm II 1k is one.
The unconditional bases are characterized as follows.
Theorem 1.51. For a basis in a Banach space X the following
rnents are equivalent.
(1) The basis is unconditional.
(2) For any permutation it of N the sequence is a basis.
(3) If for some sequence of scalars the series con-
verges in norm, then for any strictly increasing sequence of
natural numbers the series is norm convergent.
(4) If for some sequence of scalars the series con-
verges in norm and another sequence of scalars satisfies
I
for each n, then is norm convergent.
Proofs (1) (2) If (1) is true, then for any permutation 'it of N the
sequence is clearly an unconditional basis.
For the converse, assume that (2) is true and fix a permutation 'it of N.
Also, let and be the sequences of biorthogonal functionals with
respect to the bases and respectively. Since agrees with
on the vector subspace V generated by {u1, u2,. .} and V is norm dense
.
S= = = =
=
Therefore, if the series converges, then the series
converges too. Hence, in view of Theorem 1.51, is an unconditional
basic sequence.
For a special connection between unconditional bases and Banach lat-
tices see Exercise 2 in Section
Theorem 1.51 has a useful consequence regarding projections. If is
an unconditional basis in a Banach space X, then every non-empty subset
A of N gives rise to a bounded projection PA: X X defined by
PAX =
nEA
for each x = e X. If A = {1, 2,. n}, then PA is simply the
. ,
Exercises
1. Prove part (a) of Theorem 1.37
2. Prove Theorem 1.41.
3. This exercise describes a basis for the Banach spaces 1] for I p <
CX). Consider the sequence of functions defined by h1(t) = I for each
t E [0, 1] and
( -i
i
I
) , 11
Ui - 2
(0 otherwise,
for k = 0,1,2,... and I 2/c. The sequence is known as the
fl.
91t is well known that neither C[0, 1] nor L1 [0,1] has an unconditional basis; see Corol-
lazy 11.61.
.4. Ultrapowers of Bariach Spaces 37
(3) thenAEJ.
An immediate consequence of the identity A U AC = E and the
above result is the following.
Corollary 1.56. If is an ultrafilter, then for any subset A of either
AE or (exclusively) AC E
ACisafiniteset}.
A straightforward verification shows that is a filtercalled the Frchet
filter on & It turns out that every free ultrafilter includes the Frchet filter
I". To see this, let U be a free ultrafilter and let F be a finite subset of
If F = UXEF{x} E U, then it follows from Lemma 1.55 that for some x E F
we have {x} E U. This implies U = U a contradiction.
Consequently, c U for each free ultrafilter U.
The above conclusion yields the following properties for an ultrafilter U.
No finite subset of is a member of U.
By considering the collection of all filters that include the Frchet filter
(J) and using Zorn's lemma, we can easily conclude the following.
Lemma 1.57. Every infinite set admits a free ultrafilter.
For the rest of the discussion in this section U will denote a free ultrafilter
on N. By means of the ultrafilter U we now introduce a new notion of
SC(lUential convergence in a topological space.
Definition 1.58. A sequence in a topological space converges to
some x E along the ultrafilter U, in symbols, x x = limu
x the set {n E N: E V} belongs to U.
Thc point x is called a U-limit of the sequence
nEN
Proof. Since for each u = (Ui, u2,...) the sequence of real num-
bers is bounded (and so it lies in a compact subset of R), it follows
from Lemma 1.59(4) that limu exists in JR.
Now fix x (x1,x2,...) and put a = limu Let y [x].
From the fact that preserve inequalities
(see Exercise 5 at the end of the section), and y x it follows that
= [txnii] =
This implies limu IIyn $ limu IIxri By the symmetry of the situation,
limu limu also true, and so limu
is = limu a for
each y [x]. From we also get a = limu ttYLloo. Since
y {x] is arbitrary, we conclude that a
For the reverse inequality fix f > 0. It follows that the (infinite) set
B = {n N: a < < a + f} belongs to U. Now consider the
sequence y = ,c,(X) defined by Yn = if ri B and = 0
if n B. From the inclusion
Bc{nEN:
we see that {n N: = o} U. This implies x y and so
y {x]. Now notice that
Since f> 0 is arbitrary, we get [x] a. Thus, {x} = a, and the proof is
finished.
= lirri =
This shows that is a bounded operator and that ITuII Now if
x E X satisfies <1, then
IITxII = J[(Tx,Tx,. . = ITuIf f[(x,x,. . . 1[TuJI
i'toof. This conclusion follows immediately from Theorems 1.64 and 1.65
the (easily checked) fact that if X Y Z is a scheme of bounded
between Banach spaces, then = Sj,,'Tu.
Notice also that ('x)u = This implies that if T (X) is an
j
operator, then E is likewise an invertible operator and
(TuY'. D
Exercises
= f x*(f(A)) dA.
46 1. Odds and Ends
Using the definition of the integral, one can prove Cauch3r's classical
theorem in a similar manner. We state it below for future reference.
Theorem 1.70 (Cauchy's Integral Theorem). Let 0 be a bounded open sub-
set of C whose boundary 30 consists of a finite number of rectifiable simple
curves, say 30 = Assume also that each curve is oriented in
the positive directionin the usual sense of the theory of complex variables.
If U is an open subset of C such that (9 C U and f: U X is an
analytic function, then
ff(A)dA= 0.
Following the standard terminology, we say that a function f: S +
where S is a subset of C, has a power series expansion at some interior
point A0 E S whenever there exist vectors ao, ai,... in X and some positive
number r> 0 satisfying Dr(A0) c S and
f(A)
=
A where (as always) the convergence of the series is
assumed to be in the norm of X.
Lemma 1.71. A function can have at most one power series expansion at
any given point.
A E C S. If then
Ao)nx*(am) =
The following results can also be proven exactly as in the standard case.
Theorem 1.72. A function f: X, where (9 is an open subset of C, is
analytic if and only if f has a power series expansion at every point in 0
1.5. Vector-valued Functions 47
The above power series converges at least for each A E D(Ao, r),
where r = d(Ao, the distance from A0 to QC = C \ (9.
In the same manner, we can establish the following uniqueness result for
analytic functions.
Theorem 1.75. If two analytic functions f, g: 0 X coincide on a subset
of 0 which has an accumulation point in 0, then f(A) = g(A) for each
AEO.
Let us say that a function f: 0 X is weakly analytic whenever
the (standard) function x* o f: 0 C is analytic for each E Every
analytic function is obviously weakly analytic. Remarkably, as the next
theorem shows, the converse is also true.
Theorem 1.76. A function f: 0 X defined on an open subset 0 of the
complex plane C is analytic if and only if it is weakly analytic.
48 1. Odds and Ends
72
- ,
AU
x *( \
= -
d
Ic
where the circle C is considered positively oriented. This implies
X
*1
XTTi)
1 1 r 1 1
d
2ir Ic sEAm
i d
2ir
Ic
< . IAnAmI
2ir 2r
lvi
Therefore,
= sup 1x*(xn Am!,
IIx*II1
lim
7j_4QQ
=y X.
Moreover, x*(x) = x*(y) = (x* o f)'(Ao) for each
x =y and that
ff(Ao)
lim
AA0 =x
i.e., f is differentiable at A0. Since A0 is arbitrary, f is analytic.
1.5. Vector-valued Functions 49
= +
f(A) -
A E A, where the convergence is again assumed to be in the
norm.
Theorem 1.78 (Laurent). Let f: (9 X be an analytic function defined
on an open set and let ARI,R2 (Ao) (9 be an open annulus (where
need not necessarily lie in (9). Then f has a unique Laurent series ex-
pansion on AR1 ,R2 (Ao). That is, there exist (uniquely determined) vectors
c_2, c1, cO, ci, c2,... in X such that
f(A) = - (*)
1. Odds and Ends
Exercises
1. Establish the following "product rule" for derivatives: Consider two func-
tions g: 0 C and f: 0 X (where 0 is an open subset of C
and X is a complex Banach space). If both functions f and g are
differentiable at some A0 e 0, then show that the X-valued function
A [gf](A) = g(A)f(A) is also differentiable at A0 and
= fA(A)xdA.
Therefore,
x,f A(A)dA)
= [f A(A)dA](x) =
4. This is a generalization of Theorem 1.76. Let X be a complex Banach
space and let C be a (not necessarily closed) forming subspace of X*,
that is, there exists a constant 'y > 0 such that
E C and IX*II 1} 711X11
for each S X.
Show that a continuous function f: 0 X defined on an open subset
O of the complex plane C is analytic if and only if the scalar-valued
function o f is analytic for each C.
5. For a continuous function A: 0 * (X, Y), where 0 is an open subset
of C, show that the following statements are equivalent.
(a) The function A A(,\) is analytic on (9.
(b) For each x E X the Y-valued function q52,(,\) = A(A)x is ana-
lytic on (2
(c) For every x X and E Y* the standard complex-valued function
(A) = (A(A)x, = (A(A)x)
is analytic on (2
6. For each r > 0 let = {A C: IAI > r}. Assume that X is a complex
Banach space and for some r > 0 Ar + X is analytic
a function f:
having the Laurent series expansion f(A) = on the open
annulus Show that f(A) 0 if and only if = 0 for each
n = 0,1,2
7. Assume that two points a and b in a topological space X are joined by a
curve C. If C is covered by a collection V of open sets, then show that
there exist (not necessarily distinct) sets Vi, V2,... , V such that:
(a) a V1 and b E
(1)
{x1,k} is a subsequence of
'2The reader should be warned that this definition of convergence in measure is less restrictive
than the standard definition, where a sequence {x8} in Lo(ir) converges to x E Lo(ir) in measure
if for each E > 0 we have ir({w E 2 }) = 0. In practice, only sets of
finite measure are important and this is why we restrict our definition to sets of finite measure.
- - ---
1. Odds and Ends
. : k= 1, 2, . . .} is a subsequence of : k 1, 2, . .
Xn,/c on
If we let Xfl,n (the diagonal sequence), then is a subsequence of
{ that converges to zero on ft
(2) Fix some A E E with 7r(A) < oo, and assume contrary to
(3)
our claim that IA d7r 74 0. Then there exist a subsequence of
and some 0 such that IA d7r 2 for each n. The validity
of (2) guarantees that there exists a subsequence of such that
0. Since 0 and 0 < < 1 on A, it follows from the
Lebesgue Dominated Convergence Theorem that IA d7r 0. However,
this contradicts c for each n and the validity of (3) has been
established.
(3) (1) Assume that some A E E <oo and let > 0.
satisfies 7r(A)
Also, let = {w E A: c}. From the fact that t c if and only
if it follows that An {w E A: }. This implies
and consequently
= f f 0.
The next few results contain some basic properties of sequences in LQ(7r).
These results as well as some additional properties can be found in Section 71
of the book by W. A. J. Luxemburg and A. C. Zaanen [225], or in Chapter VI
of the book by B. Z. Vulikh [328].
Lemma 1.84. If cr-finite, then there exists a sequence
Lo(ir) and ir is
= U Ak)
m=Ik=rn k=n+r
= U
k=n+r k=n+r
\
k=n+r
As r 1 is arbitrary, \ A) = 0, and this guarantees that A) = 0.
\
Therefore, if we let x(w) = n 1}, then x E Lo(7r) and
O<Anxn<xforeachn.
Moreover, if Xn,kl x holds for all ri and k, then we can take the
sequence {Yn} to satisfy Yn x for each ri.
Proof. By Corollary 1.86, for each k there exists some un E such that
the sequence {xn,k: k 1} is relatively uniformly convergent to zero with
respect to the regulator un. Now, by Lemma 1.84, there exists a sequence
{ )'tn } of positive scalars and some u E that )'tnun u for each ri.
such
Put Yn = and note that for each fixed natural number ri there exists an
integer kn such that lxn,kI < < for all k kn. An easy inductive
argument shows that can be taken to be strictly increasing. If x
holds for all ri and k, then let Yn = x A (ku).
A Riesz space satisfying this property is referred to as a Riesz space with the diagonal
property.
Fundamentals of Measure Theory 57
(P) If D ,j.
0 inR, then 0 holds for 'Tr-almost
allwEft
.j,
0) for each w E B. By Egorov's theorem, we can assume that
converges uniformly to zero on B. This implies that there exists a strictly
increasing sequence such that
of natural numbers for all
w E B and all n. That is, (w) > 1 holds for all w E B and all n.
However, the latter conclusion contradicts property (P), according to which
0. This contradiction shows that y E Lo(ir). To finish the proof
note that y is an upper bound of D; in fact, y = sup D in Lo(ir).
58 - 1. Odds arid Ends
Corollary 1.89. For an operator T: E f Lo(ir'), where ir and ir' are two
cr-finite measures and E is an ideal in Lo(ir), we have the following.
( a) If T is cr-order continuous, then T is regular.
(b) If for each order bounded sequence in E that satisfies 0
we have T is a-order continuousand hence it is
a regular order continuous operator.
Proof. (a) Fix x E E+. For any sequence [0, x} and any sequence
of scalars with j 0, we have 0 A72x j 0. This shows that
0 in E, and so by the a-order continuity of T we get 0.
From Corollary 1.86, we obtain and hence, by Theorem 1.88,
0,
the set T[0, x] is order bounded in Lo(ir'). Therefore, in view of Theo-
rem 1.16, T is a regular operator.
(b) Let 0 in E. This implies that is an order bounded se-
quence and 0. By our hypothesis, 0. Now a glance at Corol-
lary 1.86 shows that T
a is a-finite, then (as we saw in Theo-
rem 1.80) the Riesz space is Dedekind complete, and hence the ideals
in L0 (7r) are Dedekind complete Riesz spaces in their own right. In particu-
lar, the classical (0 < p oo) are all Dedekind complete. For
a proof of the following important density property of order ideals in L0 (ir)
see Exercise 13 at the end of the section.
Lemma 1.90. If is a
>, ir) measure space and J is an order
dense ideal14 in Lo(ir), then for each function x E Lo(ir) there exists a
sequence JflLi(ir) such that IxnI IxI for each n and x(w)
for ir-almost all w E IL
If, in addition, x then the above sequence can be chosen
in fl Li(ir) and increasing, i.e., satisfying 0 x in Lo(ir).
Supp x = Supp = = n
n=1 n=lk=1 n=lk=lrn=1
Since fl and E E, we see that E E for all n,
k, and m. Thus, we can write Supp x = U=1 with E E for each n.
Now if V1 = B1 and = \ for n 1,2,..., then E J
and <oc for each n, Vk fl = 0 fork n, and Suppx =
The validity of the last statement of the corollary follows from the fact
that 1 E L0(7r).
Theorem 1.92. For a cr-finite measure space (12, >, it) and an ideal E in
Lo(7r) we have the following:
(a) There exists a smallest (with respect to rr-a.e. inclusion) set CE in
> such that every x E E vanishes rra.e. on 12 \ CE.
(b) If E is a band, then
E={XEL0(ir): x=0 rr-a.e. on
(c) Two bands B1 and B2 in Lo(71) coincide if and only if C21 = C22
7r (CC
\ A) = 0, and from this we get CC C A Hence, CE = AC c C,
and the proof of this case is finished.
CASE II: IT is a measure.
Fix a disjoint sequence E with = and < 00
for each n. Also, let E E} and = f E}.
Then each can be considered as an ideal in it). Therefore, from
Case I, for each m there exists a unique minimal measurable set
such that each function x E vanishes outside of the set
CE = then a direct verification shows that this measurable
set CE satisfies the desired properties. -
Assume that E is a band, and let
(b)
B = {x E Lo(it): x = 0 it-a.e. on c2\CE}.
Clearly, B is a band in Lo(7r) and .E c B. Now let 0 < f B and
consider the set C = {w E f(w) > 0}, and note that -w(C) > 0 and
C CE. From the definition of the set CE, it follows that there exists
some g E such that g 0 it-a.e. on C. This means that the measurable
set C = {w E C: g(w) > 0} satisfies it(C) > 0. Moreover, the sequence
{f A ng} c E satisfies f A ng I I xc. Therefore, f xc E .E, and thus
0 < fxc < f. This shows that E is order dense in the band B. Since E is
a band, E = B. (Notice also that E can be identified with Lo(CE,IT).)
(c) This follows immediately from part (b).
Definition 1.93. If E
is an ideal in Lo(it), then the it-a.e. unique measur-
able subset CE determined in part (a) of Theorem 1.92 is called the carrier
(or the support) of the ideal E.
It should be noticed that the order dense ideals in L0(i'r) are precisely
the ideals whose carriers coincide (it-a.e.) with
Now let E be an ideal in Lo(it). Since the measurable sets of the measure
space (CE, ECE, ii), where ECE = {A n CE: A E E}, are precisely the
restrictions of the measurable subsets of to CE (see [32, Problem 15.7,
p. 125]), it follows that the band generated by E in Lo(it) can be identified
with LO(CE, it) = LO(CE). In other words, we have the following result.
Lemma 1.94. If E is an ideal in Lo(it) and is its carrier, then E is
order dense in LO(CE).
We proceed with a brief discussion of Carathodory's classical extension
of measures defined on semirings. Recall that a collection F of subsets of a
set X is said to be a semiring if:
(1)
1.6. Fundamentals of Measure Theory 61
with the convention that inf 0 = oo. Recall that a subset E of X is said to
be rn-measurable if
rn*(A) = rn*(AflE)+rn*(AflEC)
for each A c X. It is well known that the collection
{A E 2X: A is rn-measurable}
is a cr-algebra. The relationships between this cr-algebra and the initial
semiring F as well as between rn* and rn are summarized in the next classical
theorem due to C. Carathodory; for details see [31, Chapter 3].
Theorem 1.95 (Carathodory). Let rn: F [0, oo] be a measure on a
semiring. Then the Carat hAodory extension m* of rn restricted to is
a measure that extends rn, i.e., rn*(A) = rn(A) for each A E F and every
member of F is measurable, i.e., F Moreover, we have the following
three important properties.
(a) The measure rn*: >m f [0,oo] is com4plete, i.e., rn*(A) = 0 implies
f f f(s, t)
= f [f f(s, t) dy(s)]
The meaning of the iterated integral f7 f(s, t) is analogous.
That is.
f f f(s, t) =
f [f f(s, t) dv(t)] dy(s).
1.6. Fundamentals of Measure Theory 63
fSxT
f x v)
= fTSf f(s, t) = ffST f(s, t)
fSxT
The
I xv)
= ff
TS = ffST
next result presents some basic measurability properties of ji x v-
measurable functions.
Lemma 1.99, Assume that and are two u-finite mea-
sure spaces and let f: $ x T be a x v-measurable function. Then:
(1) For v-almost all t T the function f(., t) is
(2) For all s E S the function f(s,.) is v-measurable.
1ff: S x is jointly measurable, then f(.,t) is Y1-measurable for
each t c T and f(s,.) is v-measurable for each s E S.
For a proof of the next result see [32, Problem 20.16]. See also Exer-
cise 16 at the end of this section.
Lemma 1.100. If (S, p) and (T, v) are two arbitrary rnea-
sure spaces and f: S x T R is a ,a x vmeasurable function, then there
exists a -measurable function 9: SxT R such that f = g ,axv-a.e.
If f does not have any essential upper bounds, thenas usualwe let
ess sup f = inf 0 = cc. The function f is essentially bounded above
(or simply bounded above) if it has an essential upper bound (or, equiv-
alently, if ess sup f < cc). The essential infimum of f (denoted ess inf f or
ess f(w)) is defined analogously. We have ess inf f = ess sup ( f).
The essential sup norm (or simply the sup norm) of a function
denoted by
[oo, 00] is and is the essential supremum of
the function If I. That is,
If = esssup (f I = inf{M 0: M is an essential bound of
Lemma 1.101. The essential sup norm satisfies the following properties.
Exercises
Let E be a Riesz space with the countable sup property, and let D be a
subset of E such that the supremum u = sup D exists. Does there exist
a countable subset D1 of D such that for each x E D there is an element
a E D1 such that a x? (Any subset D1 of D with this property is called
cofinal in D.)
2. Let E be an Archimedean Riesz space satisfying the countable sup prop-
erty and having a weak unit. Show that for each non-empty subset D of
E there exists a countable subset D0 of D such that D0 has the same set
of upper bounds as D. That is, whenever u E E satisfies u x for each
XE Do,thenuxforeachxED.
3. Let iv) be a measure space, and let of all subsets
be the collection
of that can be written in the form where A E and M c N
for some N E with iv(N) = 0. (As always, = (A\ M) U (M\ A)
is the symmetric difference of the sets A and M.) Verify that:
(a) is a cr-algebra and c
(b) The function p [0, oc], defined by = iv(A), is well
defined and extends IV from to a measure on
(c) If some A E satisfies 7r* (A) < oc, then A E In particular,
is a complete measure, that is, if some subset A of ci satisfies
= 0, then AE
(d) If IV is cr-finite, then =
4. Let J be an ideal in some Lo(iv)-space and let be an order bounded
sequence in J. Show that x in J if and only if p u(w) for
p(x) =
p(x) =
f
Show that for a sequence {x71} Lo(ir) and a function x E Lu(ir) the
following statements are equivalent.
(a)
(b) For each k we have Em f dir = 0.
(c) lim p(x71 x) = 0.
ri*oo
13. Prove Lemma 1.90. [HINT: Note that L1 (iv) is an order dense ideal in
Lo(ir) and that the intersection of two order dense ideals is an order dense
ideal.
14. Prove Lemma 1.96.
15. Let F be a semiring of subsets of set and let r p [0,oo) be two
set functions such that 0(A) holds for each A e r. If 0 is finitely
additive and e is a measure (i.e., countably additive), then show that 0 is
also a measure.
16. Establish the following property of the product measure: If the measures
,i and v are a-finite, then for each set C E and for each E> 0 there
exists a set of the form D = x where E E3 and E E2
for each n, such that x <E.
17. If / E Lo(iv), then show that
max{ess sup /, ess sup (/)} = ess sup I/I =
18. Assume that (11, E, iv) is a measure space and let Lt(iv). Also
let x = be the ir-a.e. pointwise supremum of the sequence
Show that x L1 (iv) if and only if there exists a pairwise disjoint sequence
{ATh} E such that div = 00.
Chapter 2
This chapter presents the general background needed to study (linear) op-
erators. It gives a concise presentation of the basic structural properties of
Banach spaces and Banach lattices and discusses the fundamental properties
of operators between these spaces. We pay special attention to the structure
of Banach lattices and operators acting on them. Banach lattices are
duced as special cases of ordered vector spaces. We take this opportunity
and review the fundamental properties of compact and weakly compact op-
erators between Banach spaces. We also state the basic compactness results
on operators dominated by compact or weakly compact operators.
We assume the reader is familiar with a basic course in functional anal-
ysis that includes an elementary introduction to operator theory.
IITxII
- TXmIf = - Xm)jI
we see that is a norm Cauchy sequence. If v x, then p Tx,
and so y = Tx, proving that T has a closed range.
For the converse, assume that T has a closed range and is By
the Open Mapping Theorem, the operator T: X R(T) has a continuous
2. 1 . Bounded Below Operators 71
inverse. So, for some c > 0 we have for all y e R(T). Letting
y=Tx,weget forallxEX.
Corollary 2.6. If a bounded operator T: X Y between Banach spaces
has a closed range and a finite dimensional null space, then every operator
S E (X, Y) close to T has also closed range and a finite dimensional null
space such that dim N(S) dim N(T).
In particular, the collection of all bounded operators in (X, Y) having a
finite dimensional null space and a closed range is an open subset of (X, Y).
JSxlI SxM =
This shows that the operator S is bounded below on V, and so its
range S(V) is a closed subspace of Y and S is olie-to-one. Now note that
S(X) = 8(N(T)) + 8(V) and that 8(N(T)) is a finite dimensional vector
subspace. Since 8(V) is closed, it follows that 8 has a closed range; see
Exercise 1 in Section 3.4. Finally, a look at Exercise 6 at the end of the
section shows that dim N(S) <dim N(T).
Proof. The implications (1) (2) (3) are obvious, and (3) (2)
follows from Theorem 2.5.
IITXnTXmIJ IETXnyM+IIYTXmIJ
= + IITrnXrn TXmJI
< [lIT + IT TmM] >0.
Let x in X and so Tx. On the other hand, from
- = - TnxnlI - . 0,
it follows that Tx = lim = y, as desired.
The next result (which can also be found in [170] and [19]) asserts that
the set of all surjective operators is an open set. It should be noticed that
its proof resembles the proof of the Open Mapping Theorem.
Theorem 2.10. If X and Y are Banach spaces, then the set of all surjective
operators from X to Y is an open subset of (X, Y).
Proof. Let T E (X, Y) be a surjective operator. By the Open Mapping
Theorem we know that T is an open mapping. Therefore, there exists some
0 <c < I such that 2cUy or
cUy C (*)
2.1. Bounded Below Operators
yflSxfl_i, and
for n = 1,2,....
Since X is a Banach space, the series is norm convergent in X,
say x = Moreover, from = 0, we see that
00 00 00
Corollary 2.12. If X and Y are Banach spaces, then the set of all surjective
and non-invertible operators fromX toY is an open-subset-of (X,Y-).- -
IIyIkr' = I J
[z]
(d) The operator T has closed range (i.e., R(T) is Il-closed) if and
only if the norms . and are equivalent on R(T).
I
111T-contznuous. In particular, every operator S E (Y) that leaves
R(T) invariant is ItT-continuous on R(T).
21. Bounded Below Operators 75
(3) For any sequence {yn} in R(T) satisfying lynII + 0, there exists a
sequence in X such that 0 and = for each m.
Proof. (1) (2) Since R(T) is closed, the norms and are equiv-
alent on R(T). That is, there exists some 5 > 0 such that
76 2. Basic Operator Theory
have [z] JJ where [z] = z + N(T). Now notice that there exists some
x E X with y = Tx and lxii (1 +
Obvious.
(3) Assume that a sequence {YrJ R(T) satisfies
(1) f 0.
Pick a sequence in X satisfying n and * 0.
From the inequality
= JlxniI
we get IYn fIT + 0. So, the identity operator I: (R(T), II. f (R(T),
is continuous. By part (d) of Theorem 2.14, R(T) is closed.
Another simplebut very usefulcharacterization of the closedness of
the range of a bounded operator is the following.
Theorem 2.16. A bounded operator T: X * Y between two Banach spaces
has a closed range if and only if there a closed subspace Z of Y such
that R(T) n Z = {0} and the vector subspace R(T) Z of Y is closed.
Proof. If R(T) is closed, then the closed subspace Z = {0} of Y satisfies
the stated properties. For the converse, assume that there exists a closed
subspace Z of Y such that R(T) n Z = {0} and the vector subspace V =
R(T) Z of Y is closed. Clearly, the vector space VIZ equipped with its
quotient norm is a Banach space. Moreover, the mapping J: R(T) V/Z,
defined by J(y) = y + Z, is a surjective linear isomorphism.
Since fJyIJ holds for each y E R(T), it follows that
the operator J: (R(T), . * (VIZ, II) is continuous, and hence it
is an isomorphism. In particular, there exists a constant c > 0 satisfying
for all y E R(T). This implies
c(IylIT
for all y E Y. That is, the norms . I! and H lIT are equivalent on R(T).
Therefore R(T) is a closed subspace of Y.
Corollary 2.17. Let T: X f Y be a bounded operator between two Banach
spaces. If the quotient vector space Y/R(T) is finite dimensional, then the
range of T is closed.
Proof. Assume first that R(T) is closed. Since the operator T: X R(T)
is a surjective operator between Banach spaces, there exists (by virtue of
the Open Mapping Theorem) some constant c > 0 such that CUR C
where UR denotes the open unit ball of R(T).
Let R(T) satisfy ft Pick a sequence satisfying
= for each n, and let be the restriction of to R(T). From
YEUR
For the converse, assume that R(T*) is closed. Put Z = R(T) and
consider the operator S: X Z defined by Sx = Tx for each x E X. Let
y* be an arbitrary linear extension of to all of Y. Then we have
S*z*(x) = z*(Tx) = y*(Tx) = T*y*(x) for all x E X, i.e., S*z* = T*y*.
This implies that R(S*) = R(T*). In other words, replacing T by S if
necessary, we can suppose without loss of generality that the range of T is
also dense. Under this extra assumption, we must show that R(T) = Y.
Since R(T) = Y, it easily follows (see Exercise 10 at the end of the
section) that T* is one-to-one. Since R(T*) is assumed to be closed, by
Theorem 2.5, the operator T* is bounded below. That is, there exists some
constant C > 0 satisfying CJIy* tIT*y* 1 for all E Next, we claim
that we have {y E Y: < C} If this is done, then (as in the
proof of the Open Mapping Y: C T(U),
from which it follows that R(T) = Y.
To establish the inclusion {y E Y: jyfj <C} pick any E Y
satisfying < C and assume by way of contradiction that y
Then, by the Separation Theorem, there exists some E such that
Exercises
1. For a bounded below continuous operator T: X Y between normed
spaces establish the following:
(a) If X is not a Banach space, then T need not have a closed rangeS
(b) The unique continuous linear extension of T to the norm completion
of X is a'so bounded below.
[HINT: Let Y be an arbitrary Banach space, X a non-closed subspace of
Y and define the operator T: X Y by Tx = x for each x E X. Then
R(T) = X is not closed in Y.]
2. Prove Lemma 2.3.
3. Let T: X p Y be a bounded, one-to-one, and surjective operator betwee
two Banach spaces. Show that T is an isornetry if and only if T and T'
are both contractions. (Recall that a bounded operator between two
normed spaces is called a contraction if its norm is no more than one.)
4. Let T E (X) be bounded below but not invertible. Show that there
exists some > such that the operator A T
0 is not surjective whenever
AJ <. [HINT: Use Theorem
5. This is a generalization of the previous exercise. Assume that X and
Y are Banach spaces and let T E (X, Y) be below but not
invertible. Then show that there exists some such that the operator
T S is not surjective whenever S E (X, Y) satisfies <E. That is,
show that the set of all bounded below and non-surjective operators from
X to Y is an open subset of (X, Y).
6. Establish the following property that was used in the proof of Corol-
lary 2.6. Assume that X = V W, where V is finite dimensional. If Y
is an arbitrary vector space, then any operator T: X Y that is one-to-
one on W satisfies dimN(T) <dim V.
7. Show that the assumption dim N(T) <oc in the last conclusion of Cor-
ollary 2.6 cannot be replaced by the assumption that N(T) is comple-
mented.
(b) If the adjoint operator T* is one-to-one and has closed range, then
show that T is surjective. [HINT: Use the preceding part in connec-
tion with Theorem 2.18.]
(c) If T carries norm bounded closed sets onto closed sets, then T
has a closed range. Moreover, all powers of T a'so have closed
ranges. [HINT: We can assume that T is one-to-one. (Other-
wise, we can replace X by X = X/KerT and T by k * Y
defined by T[x] = Tx.) Now assume * y. If has
a norm bounded subsequence, say then y belongs to the
closed set T({zi, Z2,. .}). The case oo cannot occur.
For if 11 00, then * 0, and consequently zero belongs
Proof. We establish the validity of (1.) only. To this end, assume that
equality = holds for some k. The proof is by induction
on n. For the induction step, suppose that N(Tk) = for some
n1. then we have = = 0 or,
equivalently, TThx e N(Tk+l) = N(Tc). This implies = Tk(TThx) =0
or x E Consequently, =
turns out that if the ascent and descent of an operator are finite, then
It
belongs to N(Tm) +
Letting p = rn in (*) and q = /c in (**), we get
X = N(Tm) R(Tk). (***)
Fix u E From (* * we can write U = + U2 with the vectors
U1 N(Tm) c N(Tm+l) and U2 E R(T'). But then from (*), we get
U2 = U Ui E N(Tm+l) fl = {0} or U2 = 0. Hence, U = Ui E
and so N(Tm+l) = N(Tm). Thus, /c <rn.
Next, let v E R(Tk). Using (**) we can write v = v1 + V2 with
in N(Tm) and v2 E But then it follows from (*) that
the vector = V satisfies E R(Tc) fl N(Tm) = {0} or v1 = 0.
Consequently, v = v2 E and so R(Tk) = This implies
rn < /c. Therefore, rn = k p < 00. The closedness of follows from
Theorem 2.16.
= + =
Exercises
1. Prove part (2) of Lemma 2.19.
2. Show that every operator on a finite dimensional vector space has finite
ascent and descent.
3. Consider the operator T: R3 R3 defined by the matrix
001
T=010.
000
Find the ascent and descent of T.
4. Give examples of bounded operators T: X X on a Banach space such
that:
(i) a(T) <co and 8(T) = co.
(ii) a(T) = co and 8(T) <Co.
(iii) = co and 8(T) = Co.
J[v} = +
Tothisend,flxOx*EF*andxEE+. SinceTxeFcF**andFis
a Riesz subspace of F**, it follows from Theorem 1.16 that for each z e
we have
(*)
86 2. Basic Operator Theory
x*(c) =
(IT*Ix*,x).
From this and (**), we get (jTl*x*,x) (IT*Ix*,x).
Proof. (1) Let E be an atomless Banach lattice with order continuous norm.
Assume, by way of contradiction, that E* has an atom, say 0 < q5
So, whenever some E E* satisfies 0 q5, then there exists a constant
A 0 such that = Aq5.
The order continuity of the norm of E implies that the linear functional
q5 is order continuous, and so its null ideal = {x E E: q5(IxI) = o} is a
bandand, of course, we have E = = The assumption
q5 > 0 implies {O} (and, moreover, 0 < x E implies q5(x) > 0).
Thus, there exists some 0 <u E Cgo. We claim that u is an atom of E.
To see this, let 0 <x u and 0 y u satisfy x A y = 0. We shall
show that y = 0. Let denote the band projection of E onto the band
generated by x in E. Clearly, q5 a E E* and 0 q5 q5. So, there is
some A 0 such that = Aq5. From 0 <q5(x) = = Aq5(x), we
infer that A = 1 and so q5 = Therefore, q5(y) = = q5(O) = 0,
and by the strict positivity of q5 on its carrier we see that y = 0. Hence,
u is an atom of E which contradicts our hypothesis. Consequently, E* is an
atomless Banach lattice.
(2) Now let E* be atomless, and assume that some 0 < U E is an
atom. Then the vector subspace B= {Au: A E JR} = JRu is a projection
band of E; see Lemma 2.30. So, if A = then E = A B holds. (Both
bands A and B are, of course, closed vector subspaces.) It easily follows
that E is topologically and lattice isomorphic to the Banach lattice A JR,
and so E* is lattice isomorphic to A* JR. Since A* JR has atoms, we see
that E* likewise has atoms, which is a contradiction. Hence, E must be an
atomless Banach lattice.
Exercises
1. Let be a compact Hausdorff space. Show that the Banach lattice C(12)
has order continuous norm if and only if is a finite set. [HINT: Fix any
w E and consider the set
= {x 0 x 1, x(w) 1, and x =outside
0
<
So, if y = + 1)x] + and z = [u + then we have
o <y u, 0 < z u and y A z = 0, contradicting the fact that u is an
atom. Hence, = u.]
3. Show that C[0, 1]-and L1 [0, 1] are atomless Banach lattices.
6. Let E be a Riesz space and let 0 < f E". Show that f is an atom
in if and only if f is a lattice homomorphism, that is, if and only if
f(x Vy) = max{f(x),f(y)} for all x,y E.
7. Let S, T: E F be two regular operators between Banach lattices with
F Dedekind complete. Show that:
(a) S T in r(E, F) implies 5* in r(F*, E*).
(b) IT*I ITI*.
Also, give an example where T*f < ITJ* holds true. (Compare this with
Theorem 2.28.)
8. Show that a Banach lattice E is a KB-space if and only if E is a band in
its double dual E**.
9. (DoddsFremlin [109]) Let E and F be two Banach lattices with F having
order continuous norm. If x E+, then show that the set
B = {Tr(E, F):
E T[0, x] is norm totally bounded }
is a band in r(E,F).
if ISIT.
There is an important collection of results regarding compactness
erties of operators between Banach lattices that are dominated by compact
or weakly compact operators. We list them below and for proofs and details
we refer the reader to [30]. There is no parallel for these results on general
Banach spaces.
2. Basic Operator Theory
Exercises
1. Show that a positive operator T: E F between two Riesz spaces dom-
inates an operator 8: E F if and only if T S T.
2. Let T: E F be a positive operator between two Banach lattices. Show
that for an operator 8: E F the following statements are equivalent.
(a) T dominates S.
(b) T* dominates 5*
(c) T** dominates
[HINT: Use Exercise I above.
2.4. Compact and Weakly Compact Positive Operators 91
5. Prove properties similar to the ones listed in Exercise 3 above for weakly
compact operators.
7. Show that a subset of a normed space is norm totally bounded if and only
if it is contained in the closed convex hull of a sequence that coxiverges to
zero.
is weakly compact.
Chapter 3
Operators on and
spaces
Banach lattices possess a property that has no parallel in the general theory
of Banach spaces. It can be described by saying that: Every Bariach lattice
is locally "classical." Specifically, if E is a Banach 'attice and x is any vector
in E, then the principal ideal generated by x, i.e., the vector subspace
93
94 3. Operators on AL- and AM-spaces
As we shall see, these operators are in essence the "building blocks" for the
general theory of positive operators.
It turns out that when E is a Banach lattice, then each principal ideal has
the structure of an AM-space. For a proof of the next result see [30, p. 187].
Theorem 3.4. Let E be a Banach lattice and let be the principal ideal
generated by a vector u E E. Then under the norm
= inf{A?0: I
The next two classical results allow us to consider the vectors of AL-
and AM-spaces as functions in some familiar function spaces.
The reader should keep in mind that for a compact Hausdorif space the
Banach lattice is Dedekind complete if and only if is extreinally
disconnected or Stonean (i.e., if and only if the closure of every open
subset of is also open) . For details and more on these classical theorems
we refer the reader to [30, pp. 192195].
Corollary 3.7. Every AL-space has order continuous norm.
Lemma 3.8. If E is either a or a Dedekind complete AM-space
with unit, then there is a positive projection P: E** with range E.
Proof. If E is a KB-space, then E is a band in E** (see Exercise 8 in
Section 2.3) and the band projection from E** onto E is the desired positive
projection.
Now assume that E is a Dedekind complete AM-space with unit. By
Theorem 3.3 the unit of E is a unit of E** too, and so E majorizes1 E**.
In particular, the identity operator I: E E has a positive extension to
all of E**; see [30, Theorem 2 p. 26]. Any such extension is a positive
projection on E** with range E.
Theorem 39. Assume that either F is a Dedekind complete
with unit or E is an AL-space and F is a KB-space. Then every continuous
operator from E to F is regular, i. e., (E, F) =
(E, F) has a modulus [TI.
Proof. Assume first that F is a Dedekind complete AM-space with unit,
and take any continuous operator T: E F. Since T maps the unit ball
UE of E to a norm bounded subset of F, the set T(UE) is necessarily order
bounded, and thus T is an order bounded operator. As F is Dedekind
complete, T is a regular operator and its modulus TI exists.
Assume next that E is an a KB-space. Consider an
arbitrary T E (E, F). Then T*: F* and E* is a Dedekind complete
AM-space with unit. According to the preceding case, the operator T*
is regular. This implies that T**: E** F** is also a regular operator.
Pick two positive operators A, B: E** F** such that T** = A B. By
Lemma 3.8 there exists a positive projection P from F** onto F. It remains
to notice that PA, PB are positive operators from E** to F and the equality
T = PA PB holds on E. This shows that T is a regular operator. D
Corollary 3.10. Assume that T: E F is a continuous operator, where
either F is a Dedekind complete AM-space with unit or E is an AL-space
and F is a KB-space. Then = lTIIr, the operator norm and the
r-norm of T coincide.
Recall that a vector subspace Y of an ordered vector space Z xnajorizes Z if for every
z E Z there exists some y E Y such that z < y.
3.1. AL- and AM-spaces
Fix
Since
0 and pick some unit vector x E such that JTjx L> .
=
we can find a non-empty open subset V of such that
(lTIx)(w) > E (*)
for each w E V. The RieszKantorovich formula
ITIx=sup{Tu: x<u<x}
implies the existence of some u E [x, xl and some point wO E V such that
(ITiu)(wo) >
(Otherwise the values of the function Tix on the set V would be bounded
from above by E, contradicting (*).) This implies >
and hence 11Th > As 0 is arbitrary, it follows that 11Th
The converse inequality 11Th is always true, and consequently we
have the desired equality 11Th = 1ITJr.
-Now assume that E is an AL-space and F is a KB-space. The modulus
TI: E F exists by Theorem 3.9. Furthermore, Theorem 2.28 guarantees
that ITI* = T*f. Since E* is a Dedekind complete AMspace with unit, the
first case yields 1
= JIT*II and so
The preceding result combined with Corollary 3.7 and Theorem 2.28
yields the following.
Corollary 3.11. If E is an AL-space and F is a KB-space, then the map-
ping T T*, from .C(E, F) to (F*, E*), is a lattice isometry.
N F**
I
T**lx = E E** and <x} ITJx,
We are now ready to state and prove the analogue of Theorem 3.12 for
compact operators.
that is a KB-space. By
Theorem 3.9, the modulus TI of T exists. Next, notice that T*: F*
a compact operator and E* is a Dedekind complete AM-space with unit. By
the preceding case, T* has a compact modulus IT* I. Finally, Theorem 2.28
implies ITI* IT* , and so ITI* is a compact operator. The latter guarantees
that itself is a compact operator.
Exercises
1. A vector u in a subset A of a vector space X is called an internal point
of A if for each x E X there exists some > 0 such that u + Ax E A for
all Al A0.
Show that in an ordered vector space X the strong units are precisely
the internal points of the positive cone
2. If E is an AM-space with a unit e, then show that the formula
= inf{A>0: xI Ae} = min{A 0: JxJ Ae} (*)
9. Show that the lattice operations in an AL-space need not be weakly se-
quentially continuous.
10. Let E be a Banach lattice. For x E
E E**: x x}.
Establish the following:
(a) [x, xl and ftx, 4 are both convex sets.
(b) [x, x] is a(E, E*)_closed and is a(E**, E*)_compact.
(c) [x,x] is a(E**,E*)_dense in
[HINT: To see (c), assume that [x, x] is not a(E**, E*)_dense in ftx, x}J.
Then, by the Separation Theorem, there exists some E E* satisfying
x*(y) = 0 for each y E [x,x] and x*(y**) 0 for some E It
follows that
Ix*I(x) = sup{x*(y): yE E and x <y x} 0.
11. Show that every regular operator from an AM-space with unit to any
p-space, where 1 p < oo, is compact. [HINT: Use the fact that the
order intervals of 4-spaces with 1 p < oo are norm compact sets.]
12. Let E be an AM-space. Extend the norm of E to its Dedekind completion
as in Exercise 21 of Section 1.2. Show that the Banach lattice E'5 is
also an AM-space. If E is an AM-space with unit e, then Ea is also an
AM-space with unit e.
13. Assume that S and Q are two compact metric spaces, is an arbitrary
(finite) Borel measure on 5, and K: S x Q + is a continuous function.
Also let T: C(S) C(Q) denote the hitegral operator with kernel K,
i.e., T is defined by
= sup
f 5K(s,w)1 dy(s).
[HINTS: (a) Let d and p be the metrics on S and Q, respectively, and
let c > 0. Fix some S > 0 such that d(s1, S2) + p(wi, W2) < S implies
K(si,wi) < c. Now if x E C(S) satisfies 1 and
= lim I
TL+OO
14. Show that every Markov operator carries order units to order units.
15. Every Euclidean space can be considered either as an AM-space with
unit e = (1, 1,. .. , 1) or as an AL-space under the usual norms
and
l<i<TL
i=1
respectively. An rn x n withnon-negative real entries
is called Markov (resp. stochastic) if = 1 for each 1 < i rn
(resp. = 1 for each 1 j n).
Now let A be an m x n matrix with non-negative entries and consider
A as an operator from R7L to Show the following:
(a) The positive operator A: . II II
is a 1\'Iarkov
new norm, assume first that E C(f1), the Banach lattice of all
continuous functions on a compact Haiisdorff space ft2 An easy argument
shows that for each pair f,g C(Q) and each w E Q, we have
sup {f(w) cos 9 + g(w) sin 9] sup [f(w) cos 9 + g(w) sin 9]
OE[O,21rj OER
= + [g(w)}2.
Conseqriently,
=
where the supremum is taken in in the usual lattice sense.
We now turn our attention to the general case of a real Banach lattice
B. Let x, y E E and consider the ideal generated in E by the element
u = lxi + By Theorems and 3.6, we know that is lattice isometric
to a This implies (in view of the preceding discussion) that the
supremum
sup [x cos 9 + y sin 9] = sup [x cos 9 + y sin 9]
OE[O,2ir] OER
exists in (and hence in E). The above sripremrim is called the modulus
of the element z = x+zy E That is, the modulus of an arbitrary element
z = x + zy E is the positive element izl of B defined via the formula
= sup [x cos 9 + y sin 9].
OE[O,21r1
Notice that
lzt= xzyj = = (=
It can be shown that for arbitrary z, z1, z2 E and A C, we have:
(1) zI 0 and IzI=0 if and only ifz=O;
(2) = Al . Izi; and
(3) < + IZ2I.
21f we want to emphasize that we deal with real-valued functions rather than with complex-
valued functions, then we shall denote by CR(cl). The symbol Cc(cl) will denote the complex
Banach space of all complex-valued continuous functions on ci equipped with the supremum norm.
Complex Banach Lattices 105
then the above properties (1), (2), and (3) guarantee that is, in fact,
a norm on Now notice that if z = x + iy E then the inequalities
xI Izi, yI and Izl jxf + y[ imply that
+ lylf) lxii +
Therefore, not only is Ic a norm on but it is also equivalent to the
standard norm on defined by
I
z = sup lix cos 0 + y sin 0
OE[O,2ir)
The reader should also observe that = jIziI for all z E E and Izil
implies z1 Ic
From now on, we will always assume that on the complexification
E the norm is defined by (*). Specifically, we have the
following definition.
The rest of this section will demonstrate the advantages of this norm. A
linear isometry T: between two complex Banach lattices is said to
be a lattice isometry if ITzI = for each z E Two complex Banach
lattices = B zE and F zF are said to be lattice isometric if
there exists a surjective linear isometry T: satisfying = TIzI
for all z E We have the following.
11
1 = IIzII
From Lemma 3.18, it should be clear that the complex version of the
KakutaniBohnenblustKrein Representation Theorem 3.6 can be stated as
follows.
Theorem 3.20. If E iE is a complex AM-space with unit e, then
there exists a compact Hausdorff space (which is uniquely determined up
to homeomorphism) such that is lattice isometric to the complex AM-
space = and e corresponds to the constant function I
3To avoid unnecessary notation, we shall use the symbol T instead of whenever this does
not cause any ambiguity.
3.2. Complex Banach Lattices 107
for z = x + zy E
all We shall say that an operator T + iS Fc) is
regular if both operators T, S (E, F) are regular. The collection of all
(complex) regular operators will be denoted 4(Ec, Fc). Clearly, the vector
space 4(Ec, Fc) coincides with the complexification of 4(E, F), that is,
4(Ec, Fc) = F) F).
From now on, let us assume that F is also Dedekind complete. Then,
by the RieszKantorovich Theorem 1.16, 4(E, F) is a Dedekind complete
Riesz space which is a Banach lattice when equipped with the r-norm; see
Theorem 1.32. In particular, 4(Ec, F) F) is a
plex Banach lattice. Moreover, every regular operator T + iS has a modulus
IT + zSJ in Fe). Remarkably, the modulus JT + zSI is given (just as in
the real case) by the RieszKantorovich formula. The details are included
in the next result.
Theorem 3.24. Let E and F be two real Banach lattices with F Dedekind
complete, and let T = T + iS E 4(Ec, Fc) = F) irr(E, F). Then
the modulus ITI of T satisfies the RieszKantorovich formula, i.e., for each
u E E+ we have
+'' = u. It is well known that sup D = TIu; see [30, Theorem 1.16,
p. 15].
108 3. Operators on AL- and AM-spaces
x cos
For the reverse inequality, we shall need the following two (easily
checked) simple facts:
(1) If w E satisfies 1w! v for some v E then TwJ +
[SJ)v.
ITzt =
T(z + +
Ru + E a ITlu
is also true. This inequality, combined with (*), establishes the validity of
the RieszKantorovich formula.
If 1(u) = sup
Exercises
1. If E is a Banach lattice and z = x + zy E then show that
(c)
Use the above properties to verify that the function [1. + R, defined
by =1 is a norm.
3. Show that the complexification of coincides with Moreover,
show that for each 1,9 we have If + = +g2 and hence
If k/f2
4. Show that the complexification of R72 coincides with C72. Also, show that
all norms on (or R72) are equivalent. [HINT: Let . be a norm on
C12 and let . denote the 1-norm, i.e., Iz Iii = An easy
argument shows that the identity mapping I: _ (C72, 1) 11 1
and max
i=1
Show that:
IlAlk and
[HINT: We shall indicate how to establish the second formula only. Let
= If < 1 and y = Az, then
= >
(a) Show that the ideals in are precise'y the vector subspaces of the
form J J0 'iJ0, where J0 is an ideal in E.
(b) A bounded operator T: F F on a (real or complex) Banach lattice
is said to be ideal irreducible if there is no non-trivial c'osed ideal
in F which is T-invariant. That is, T is ideal irreducible if and only
if T(J) J and J a closed ideal in F imply either J = {0} or J F.
If T: E E is an operator on a real Banach 'attice, then show that
T is ideal irreducible if and on'y if is likewise ideal
irreducible.
[HINT: Assume that J is an ideal of
E E with x + zy E J}.
E
we tarn into a (complex) Hilbert space. Show that the standard norm
on defined by
= + = sup IS cos0 +
OER
J\I(x,y) = +
3.Operators on AL- arid AM-spaces
subordinate to the open cover {V1,.. ,. see [31, Theorem 10.9, p. 85].
Put = and note that
- = -
(w) [z(w) - (w) E .
Clearly, each central operator is regular and every positive central oper-
ator is a lattice homomorphism. Moreover, it is easy to see that the center
2(E) is an ideal in 4(E), the vector space of all regular operators on E.
The central operators are special examples of operators known as ortho-
morphisms. An order bounded operator T: E E on a Riesz space is said
to be an orthomorphism if A = 0 implies xl A ITyl 0. Recall
also that an operator T: E E on a Riesz space is band preserving if
T(B) B for every band B of E.
3.3. The Center of a Banach Lattice 113
In spite of the fact that for an arbitrary Riesz space E, the partially
ordered vector space need not be a Riesz space, the center Z(E) is
always a Riesz space. The details are in the next result.
Theorem 3.30 (BigardKeimel [62}; ConradDiem [93]). If E is a Riesz
space, then Z(E) is a Riesz space. Its lattice operations are given by
ITI(lxI) = T(IxI)I =
for each x E E.
a proof of Theorem 3.30 see [30, pp. 107109]. In fact more is true.
For
The center of a Banach lattice is always an AM-space with unit.
Theorem 3.31 (Wickstead [334]). For a Banach lattice E the following
two statements are trae.
(1) The center 2(E) equipped with the operator norm is an AM-space
with unit. The unit is the identity operator I, so that for each
operator T E 2(E) we have
11Th = II 1
=inf{A 0 TJ
Moreover, 2(E) is a unital Banach algebra.
(2) If E is Dedekind complete, then the center 2(E) coincides (as a
Riesz space) with the band generated by the identity operator I in
4(E); and so
114 3. Operators on AL- and AM-spaces
Now let us discuss the center of an AM-space with unit. To start with,
recall that (by Theorem 3.6) an AM-space with unit e is lattice isometric to
some C(Ifl-s pace so that the unit e corresponds to the constant function 1
on ft Every function x E gives rise to a central operator (also called
a multiplication operator) defined by = xy. A
moment's thought reveals that
= sup sup
JIyIh,ol
Z(E). That is, subject to the above identification, we have the following
result. 0 (fr V i(:: , (1
complete Riesz space and an algebra. It turns out that for every Archi-
medean Riesz space E there is a unique (up to homeornorphism) Stonean
space Q and a lattice isomorphism 'iv from E into C (Q) such that for each
f
0 < E C(Q) there exists some 0 < x e E such that 0 < 'ir(x) < f, i.e.,
'ir(E) is order dense in C(Q). If E has a weak unit e, we can assume that
iv carries e to 1, i.e., 71(e) 1, the constant function one on Q. The unique
Stonean space Q is referred to as the Stone space of E. For all
purposes, we can identify E with 71(E) and consider the elements of E as
extended real-valued functions on Q. For details about the representation
of Riesz spaces we refer the reader to [225, Chapter 7]. Summarizing the
discussion above, we have the following important result.
Theorem 3.35 (MaedaOgasawaraVulikh). If E is an Archimedean Rie$z
space, then there exists a uniqueup to homeomorphismStonean space Q
and an order dense Riesz subspace E of the Dedekind complete Riesz space
C(Q) that is lattice isomorphic to E. Moreover:
(1) If E has a weak unit e, then the lattice isornorphism can be chosen
to carry e to the constant function 1 on Q.
(2) If E is Dedekind complete, then E is an order dense ideal in
C(Q).
(3) If E is Dedekind complete with a strong unit e, then E = C(Q).
It is customary to identify E with E and we shall follow this custom. The
unique Dedekind complete Riesz space C(Q) is called the MaedaOgsa-
waraVulikh completion of E. It is also known as the universal com-
pletion of the Riesz space E. For details about the universal completion
see [29, Chapter 7].
From the general theory of orthomorphisms we also have the following
characterization of the center of a Banach lattice. For a proof see [30,
Theorem 8.28, p. 121].
Theorem 3.36. If Q is the Stone space of a Banach lattice E, then
Z(E)_{feC(Q): fgeE for all gEE}.
Moreover, if E is Dedekind complete, then Z(E) = C(Q).
We point out here an important difference between a Dedekind complete
Banach lattice and a general Banach lattice. From Theorem 3.36, it follows
that the center Z(E) of a Dedekind complete Banach lattice E is always
a non-trivial AM-space, that is, 2(E) {AI: A an arbitrary scalar}.
However, a non-Dedekind complete Banach lattice can have a trivial center.
An example of such a Banach lattice was provided first by
A. Goullet de Rugy [142]; see also [336].
3.3. The Center ofa Banach Lattice 117
(Z+R)aI =
= ha+Pa(Ra)j ('yS)IjZIJa,
Exercises
1. Show that a positive operator T : E E on a Riesz space is an ortho-
morphism if and only if I + T is a lattice homomorphism.
2. Assume that E is a Dedekind complete Banach lattice. Use Theorem 1.17
to present an alternate proof of Theorem 3.30.
3. Show that every central operator T: E E on a Banach lattice leaves
the principal ideals invariant. That is, show that holds for
each u E E+.
4. Prove Theorem 3.31.
5. If is a finite measure, show (with appropriate identifications) that
= for each 1 p 00.
6. Prove the second part of the MaedaOgasawara--Vulikh Representation
Theorem 3.35 by establishing the following general result for Riesz spaces.
If F is an order dense Riesz subspace of an Archimedean Riesz space
E and F is Dedekind complete in its own right, then F is an ideal in E.
7. Show that a central operator on a Banach lattice commutes with all band
projections.
8. This exercise justifies the name "diagonal projection" for the band
tion of Lr(E) onto the center of the Banach lattice E. Let E = the
Euclidean ri-dimensional (Riesz) space. Then (E) = 4(E), the vector
space of all ri x ri matrices with real entries. If D is a diagonal matrix
with diagonal entries d1, . . . then we write D = diag (d1,.. .
Establish the following properties.
(a) If A = and .8 = are two ri x ri matrices, then A .8 holds
in 4(E) if and only if for all i and j.
(b) If A = and .8 = are two ri x ri matrices, then
= AV B = V and A A B = A
9. Prove the second part of Theorem 3.36. That is, show that if E is a
Dedekind complete Banach lattice, then Z(E) = C(Q), where Q is the
Stone space of E.
10. Recall that the principal band generated by a vector x in a Riesz
space E is the smallest band that contains x. From Theorem 1.27, we
know that = {y E E: yIAriIxI I If E is Dedekind complete, then
is always a projection band, i.e., = E. The band projection
of E onto will be denoted by
Now let E be a Dedekind complete Banach lattice and let T, Z: E E
be two disjoint regular operators (i.e., ITI A ZI = 0) with 0 3/: Z E 2(E).
Show that for each 0 < x E E and each 0 < E < there exists a
non-zero component a of x (i.e., a A (x a) = 0) such that Pa(lTIa) Ea.
[HINT: We can suppose that T and Z are positive operators with Z
satisfying 0 < Z I. Fix 0 < E < Replacing E by we can
3.4. The Predual of a Principal Ideal 119
T*J,
Figure 1
= [Tx]. (**)
To see that the value TQ([x}) in (**) is well defined, let y E [x], i.e., let
[y] = [x} or = 0. Since T* leaves invariant, there exists some
A0 satisfying T*b This implies
0< =0,
3.4. The Predual of a Principal Ideal 121
and hence TyI) = 0. This shows [Tx] = [Ty], and so (**) defines a
= =
= =
we see that E/Nj, E/Nj, is also continuous. Let denote the unique
continuous extension of T1,: E/Nj, + to all of To see that the
diagram shown in Figure 1 commutes, note that for all
x E, we have:
= = =
= x*(Tx) = (Tx,x*) = (x,T*x*), and
([x], = ([x], = T*x*(x) (x,T*x*)
morphism.
Exercises
1. Establish the following useful result: If Y and Z are closed subspaces of a
normed space X, one of which is finite dimensional, then Y+ Z is a closed
subspace of X. [HINT: Assume that Y is closed, Z is finite dimensional,
and let it: X X/Y denote the quotient map. Then ir(Z) is a finite
dimensional subspace of the normed space X/Y, and so 7r(Z) is a closed
subspace of X/Y. Now note that Y + Z ir1 (-ir(Z)).]
122 3. Operators on AL- and AM-spaces
123
124 - 4. Special Classes of Operators
in the range T(V), then there exist (uniquely determined) linear functionals
fi, , on V such that T(v) = i.e., T =
It is easy to see that an operator T: V > W between two vector spaces
is a finite-rank operator if and only if there exist (not necessarily linearly
independent) vectors UI.,. in W and linear functionals ga,. . , on V
. . , .
The validity of the converse of Theorem 4.7, i.e., of the statement "every
separable Banach space with the approximation property has a basis," was
another long standing open problem that was solved in the negative by
P. Enflo in [120]. The situation does not improve even when X and Y are
Banach lattices; see [315].
Now let us discuss the connections between the approximation property
and the topology of uniform convergence on compact sets. Fix two Banach
spaces X and Y, and let denote the collection of all non-empty norm
compact subsets of X. For each C E define the seminorm on (X, Y)
via the formula
pc(T) = max
xEC
Itshould be clear that the family of seminorms generates a Blaus-
dorif locally convex topology Y) on (X, Y). Also, it should be obvious
that a net (X, Y) satisfies T in (X, Y) if and only
if converges uniformly to T on every compact subset of X. For this
reason, the topology Y) is called the topology of uniform conver-
gence on the compact subsets of X. We write for X).
Two easy characterizations of the approximation property in terms of
the topology of uniform convergence on compact sets are included in the
next result.
4.1. Finite-rank Operators 127
Lemma 4.8. For a Banach space X the following statements are equivalent.
(1) X has the approximation property.
(2) The identity operator I: X > X belongs to the of
the vector subspace of all finite-rank operators in (X).
= =
IIYnlI] = Mpc(T).
11f > 0 holds for each and <oc, then choose a strictly increasing sequence
of natural numbers such that < Letting = 1 for 1 i <Icr and A = n > 0
for i < we see that co and <co.
128 4. Special Classes of Operators
T R =
on C,
0 and
so 1q5(T) b(R)j = R)J Mpc(T R) = 0
implies q5(T) = q5(R). This shows that is a well-defined linear functional.
From
Ty2,. . = = MII(Tyi, Ty2,. .
= Ty2,...))
=
Since is a bounded sequence, we also have . <ce, and
the proof is finished.
Proof. Assume first that X has the approximation property, and let two
sequences X* and X satisfy IIxJI . < cc and
=
Also, let 0. denote the linear functional
on (X) defined by q5(T) Then for each rank-one operator
x we have
x) = = =
= x*(0)=O.
4. 1 . Finite-rank Operators 129
For the converse, assume that the condition is satisfied, and let be
a linear functional on (X) that vanishes on the vector
subspace of all finite-rank operators. By Theorem 4.9 there exist two se-
quences X* and X with < oo such that
b(T) = holds for each T E If x is an arbitrary
rank-one operator, then we have
0= =
Proof. Assume that has the approximation property, and let two se-
quences X* and X satisfy . < oo and
= 0. Then, considering each as a vector in it follows
from Lemma 4.3 that = = 0. Since the
Banach space has the approximation property, Corollary 4.10 implies
= 0. But then the same corollary guarantees
that X has the approximation property.
A. Grothendieck has also shown that a Banach space X has the ap-
proximation property if and only if for each Banach space Y the finite-rank
operators from Y to X are dense in K(Y, X), the closed vector subspace of
compact operators. This result is stated next.
Theorem 4.12 (Grothendieck [141]). A Bartach space X has the approxi-
mation property if and only if for each Bartach space Y the vector subspace
of all finite-rank operators X is norm dense in X).
S = X X such that x
for each E Uy we have
R=ST= Y
yE1\U},
then Y is a Banach space. Its closed unit ball coincides with U, that is,
{y E Y: 1} = U. Moreover, if we let J: Y X denote the natural
embedding (that is, Jy = y for each y E Y), then from J(U) = U, it follows
that J is a compact operator. Therefore, according to our hypothesis, there
exists a finite-rank operator 0 E (Y, X) satisfying
y y E U.
The proof would be finished from (**) if we knew that each E belongs
to Unfortunately, this might not be the case. However, the proof can be
completed if we can verify that for each E and each S > 0 there exists
some E X* satisfying I y*(x) x*(x)f <S for each x e C or equivalently
IY*@1n) X*(Xfl)I <S for each n.
4.1. Finite-rank Operators 131
Exercises
1. Let f, . , be functionals on a vector space V. Show that f
linear
lies in the linear span of fi,. , if and only if
. . Kerf.
In particular, a collection of linear functionals on a vector
. ,
= = = .1
r=I v=I i=1 i=I i=1
T(x) =
Lemma 4.17 (de Pagter [260]). Let E be a (real or complex) Banach lattice
with a quasi-interior point and let lxi Then there exists a sequence
of operators on E such that 0 and each is dominated
by the identity operator.
Proof. Let u > be a quasi-interior point of E. Replacing u by u + yi,
0
we can assume that u. Now let us view as a with u
corresponding to the constant function 1 on ft So, x,y E
For each n consider the multiplication operator p generated
by the function = Le., for each z E The unique
continuous linear extension of to all of E will be denoted by again.
138 4. Special Classes of Operators
Proof. (1) The identity TMj, = means that T(bx) = bTx for each
x E C(12). By induction, we get = bmTx for each n 2 0 and all
x E C(12). Thus,
T(p(b)x) = p(b)Tx (*)
for each polynomial p(t) of one real variable t and all x E
Letting x = 1 and h = Ti, it follows from (*) that T(p(b)) = hp(b) for
each polynomial p. In other words, Tx = hx for all x in the unital algebra
generated by q5. Consequently, by continuity, Tx = hx for all x in the closed
unital algebra generated by q5.
= = =
On the other hand, since vanishes off V and = c on V, we also have
that
= = =
Therefore, K is also a positive operator.
(2) (3) (4) Obvious.
holds for all f e and for each E where as usual M: (Q) + (Q)
is the adjoint operator of Mf.
2Here we use the conclusion of Exercise 1 in Section 4.1: A linear functional f belongs to
the span of the linear fun fit.. if and only if
. Ker C Ker f.
4.2. Multiplication Operators 141
= [fl =0
Exercises
1. Prove the following implications:
Order Unit Quasi-interior Point Weak Order Unit
Also show that in general no converse implication is true.
2. Show that a Banach lattice E has a weak unit if and only if E has an
at most countable maximal subset of pairwise disjoint vectors. [HINT: If
{ x1, .} is a maximal countable set of pairwise disjoint positive unit
vectors, then x = is a weak unit]
3. Show that every separable Banach 'attice has a quasi-interior pointand
hence it also has a weak unit.
142 4. Special Classes of Operators
The important conclusion from Theorem 4.21 is the following. When one
represents as a then all vectors in J as well as all vectors TkX,
xE can be viewed as continuous functions on the compact Hausdorif
space We can then use the properties of the continuous functions to
144 -- 4. Special Classes of Operators
momorphism if and only if there exists some c> 0 and some wo E (both
uniquely determined) such that =
c 0, thenis clearly a lattice homomorphism,
that is, A g) = for all f, g e
For the converse, assume that R is a lattice homomorphism.
Clearly, is a positive linear functional. Hence, by the Riesz Representation
Theorem, there exists a unique regular Borel measure p on satisfying
ff fe We claim that the support of ,a is a
singleton. To see this, assume that two distinct points, say s and t, belong
to the support of ji. Pick two functions f, g e satisfying f A g = 0 and
f(s) = g(t) = 1. This implies
0 Ag) = = min{ffd/2,fgd/2} > 0,
which is a contradiction. Hence, the support of consists of a single point,
say wo. This implies
f =
f e C(cl).
The algebraic homomorphisms from a to the complex num-
bers are characterized in the next result.
Lattice and Algebraic Homomorphisms 145
q E Q, we have
(8q 0 T) (f) = 8q(Tf) = [Tf] (q).
Since T is a lattice homomorphism, the linear functional 8 a T: C(11) R
is a lattice homomorphism. Thus, by Lemma 4.23, there exists a unique
constant w(q) 0 and some (not necessarily unique) point T(q) E such
that
{Tf](q) = (8q o T)(q) = w(q)f(r(q)).
Letting f = lci, we obtain w = E C(Q). Moreover, if w(q) >
then 0,
T(q) is uniquely determined. We leave it as an exercise to verify that 'r is
continuous on the set {q E Q: w(q) > 0}.
result was generalized by A. Wilansky [337]. Nowadays, using the deep re-
suits of \'V. T. Cowers and B. Maurey from [137] and [138], it has become
possibie to produce easily a Banach space X such that (X) has non-trivial
multiplicative functionals; see Exercise 7 in Section 7.1 and [200]). All the
Banach spaces X with non-trivial multiplicative functionals on (X) men-
tioned above are non-reflexive. It was P. Mankiewicz [231] who constructed
a reflexive (even a superreflexive) Banach space X with infinitely many mul-
tiplicative functionals on (X). See also [97] for another construction of a
Banach space with a similar property.
liVe shall close the section with an interesting extension property of lat-.
tice homomorphisms. In order to describe this property, we need some pre-
liminary discussion regarding extensions of positive operators. Recall that a
mapping p: X Z, from a (real) vector space to an ordered vector space,
is said to be sublinear if:
(a) p(x + y) p(x) +p(y) for all x,y E X.
(b) p(Ax)=Ap(x)foraliAOandalixeX.
An almost verbatim repetition of the proof of the classical HahnBanach
theorem yields the foliowing result regarding extensions of linear operators.
Theorem 4.31 (HahnBanachKantorovich). Let Y be a vector subspace
of a (real) vector space X, let F be a Dedekind complete Riesz space, and
let p: X F be a sublinear mapping. If an operator T: Y F satisfies
Ty p(y) for all y e Y, then there exists an operator 8: X F that
extends T and satisfies Sx <p(x) for all x E X.
This implies, using (*), that SxI = for each x E E, that is, S is a
lattice homomorphism.
Exercises
and F = C(Q) and that T is a Markov operator. Now use Theorem 4.28
and view T: C(Q) as a composition operator.]
11. Prove Theorem 1.35. [HINT: Use the preceding two exercises by observing
that if T: E F is a bounded operator between two Banach lattices,
then T*y* aT for each E F*.]
12. Let E = and let c denote the Riesz subspace of all convergent se-
quences. Show that the limit functional Lim: c p defined by
im(x) = x= (x1, x2,...) E c, is a lattice ho-
momorphism that has a lattice homomorphic linear extension to
31f W is a vector subspace of a vector space V, then the codirnension of W is the dimension
of the quotient vector space V/I'V If T: V U is a linear operator between vector spaces, then
the rank of T is the dimension of the range R(T) and the corank of T is the codimension of the
range R(T).
156 4. Special Classes of Operators
Proof. The fact that the operator T has closed range follows from Corol-
lary 2.17. For the index formula notice that according to Theorem 2.13 we
have d(T) = dim [Y/R(T)] = dim N(T*).
In our discussion we shall use the following lemma whose proof follows
from Exercise 6 in Section 4.1. Many authors use the conclusion of this
result as the definition of a Fredhoim operator.
Lemma 4.39. If T: X Y is a Fredholm operator between Bamach spaces,
then there exist a closed vector subspace V of X and a finite dimensional
vector subspace TV of Y such that
X = N(T) V and Y = R(T) W.
In particular, the surjective operator T: V R(T) is an isomorphism.
Proof. The proof is "combinatorial" in nature and is based upon the fol-
lowing simple algebraic fact:
o If W isa vector subspace of a vector space V, then there exists an
algebraic complement of W, i.e., there exists some vector subspace
U of V such that W e U = V.
(For a proof see Exercise 3 at the end of the section.)
The key object for our proof is the subspace N1 R(T) fl N(S) of Y.
Pick a subspace N2 of Y such that N(S) = N1 N2. So, we have
n(S) = dimN1 +dimN2. (*)
inclusion N(T) c N(ST), we see that rt(T) < oo. Also, from (*), we get
dim Y2 <oo, and (*-*) yields d(T) <oc. Thus, T is a Fredhoim operator.
Finally, suppose that ST and T are Fredhoim operators. Then (*-*)
yields that dim N2 <oo and dim Y2 <oo, and (* * *) implies dim N1 <oo.
But then from (*) and we obtain ri(S) <oo and d(S) <oo. That is, S
is a Fredholm operator, and the proof is finished.
In order to obtain more properties of Fredhoim operators, we need the
following useful lemma on the existence of "almost perpendicular" vectors.
Lemma 4.44 (F. Riesz). If Y is a proper closed subspace of a Banach space
X, then for each 0 < 1 there exists some unit vector x E X such that
can be proven in a similar manner. We already know that the vector space
N(I K) is finite dimensional. Since (I can be written as I K1
for some K1 E IC(X), the above conclusion shows that N((I is also
finite dimensional for each ii.
Now assume by way of contradiction that K) = oo. This means
that N ((I K)12) is a proper subspace of N((I for each ii. Conse-
quently, by Riesz's Lemma 4.44, there exists a sequence of unit vectors
in X such that E N((I for each ii and xli I for all
4.4. Fredholm Operators 161
for all n-i and n. This implies that the sequence does not have any
convergent subsequence, contradicting the compactness of K. This contra-
diction establishes that I K has a finite ascent.
(2) From the preceding conclusions we already know that N(I K*) is
finite dimensional: -Therefore,- by Theorem 2.13 we have-
d(I K) = dim [X/R(I K)] = dimN(I K*) <00.
K is a Fredholm operator.
This implies that I
To show that the index of I K equals zero, we proceed as follows.
By part (1) the operator I K has finite ascent and descentwhich by
Lemma 2.21 coincide. Let p = K) = 5(1 K). Then Theorem 2.23
guarantees that X = N((1 K)P) e R((1 K)P). This identity implies that
the Fredhoim operator satisfies
i((1 K)P) = dimN((1 K)P) dimN((1 K)P) = 0.
Now applying the Index Theorem, we obtain that i((1K)P) = p[i(1K)],
and so i(1 K) = 0. The proof is finished.
Sy {Sj'(Iy Sj'(Tv) = v.
+Tv) =
This implies that for each x = m + v E N(T) e V = X we have
STx = BT(m+v) = Sj'Tv = v = (Ix PN)(Th+V) = (Ix PN)X.
Similarly, for each y = w + Tv E TV T(V) = Y with v E V we have
TSy=Tv=(IyPw)y.
So, ST = and TS = ly and the validity of (2) has been
established.
Now a glance at the Index Theorem 4.43(2) shows that the operator T + C
is Fredhoim and satisfies
i(S)+i(T+ C) = i(S(T+C)) = + SCK1) = 0.
Thus, i(T + C) = i(S), and from (**) we get i(T + C) = i(T).
Therefore, we have established that if C E C(X, Y) satisfies MCI! < iih'
then T + C is a Fredhoim operator and i(T + C) = i(T). This shows
that Jred(X, Y) is an open subset of L(X, Y) and that the index function
i: Jred(X, Y) Z is continuous. The proof is finished.
There is a similar result regarding semi-Fredholm operators. To establish
it, we need some preliminary discussion.
Lemma 4.49. For a Banach space X we have the following.
Proof. (1) Let Y be the linear span of x1, x2, ... , We claim that X = Y.
If X Y, then by Riesz's Lemma 4.44 there exists some unit vector x E X
such that d(x,Y) = 1. In particular, we have Ix xjJJ d(x,Y) = 1 for
ball of the Banach space X/Z. From our assumption, it follows that for each
xE there exists some i such that jJ7r(x) < 1. By part (1), the
vector space X/Z is finite dimensional.
The next result deals with the codimension of an operator with a closed
range.
Lemma 4.50. For a bounded operator T: X Y between Banach spaces
with a closed range we have the following.
(1) If d(T) = 0 and N(T) is finite dimensional, then every operator
in C(X, Y) close to T also has a closed range, infinite defect, and
a finite dimensional null space.
(2) If n(T) = cc and R(T) has a finite codimension, then every oper-
ator in C(X, Y) close to T also has a closed range, infinite nullity,
and a finite codimension.
Fredhoim Operators 165
Proof. (1) The fact that for each operator S E (X, Y) close enough to T
the range R(S) is closed has been established in Corollary 2.6. The only
thing we need to verify is that d(S) = cc and dim N(S) < oo for all operators
S E (X, Y) close to T.
We assume first that T is one-to-oneS In this case, since R(T) is closed,
for some > 0 and all x E X. Pick some f E (0, 'y) such that
<1. We claim that every operator S E (X, Y) with IS <f has
a closed range, an infinite defect, and a finite dimensional null space. To
establish this, fix some S E (X, Y) satisfying HS <
From IISxII ('y that S is bounded
6)fIxfI for each x E X, we see
below on X. This implies that N(S) = {0} and that S has a closed range.
Moreover, for each y E R(S) the vector x = 5'y satisfies the estimate
(*)
Lemma 4.49(2). Consequently, the vector space Y/R(T) must be finite di-
mensional, which is impossible.
For remains to notice that the above assumption
the general case, it
that T is one-to-one is not essential. Indeed, the null space N(T) is, by
hypothesis, finite dimensional and so it is complemented in X. Let V be a
closed vector subspace of X such that X = N(T) V. Then the restriction
of T to V is one-to-one, and consequently the previous case applies to this
166 4. Special Classes of Operators
Hence, if S e
restriction. Y) is close to T, then 8(V) has an infinite
codimension in Y. Since 5(X) = S(N(T)) + 5(V) and S(N(T)) is finite
dimensional, it follows that 5(X) also has an infinite dddimension in Y.
(2) Assume that a bounded operator T: X Y between Banach spaces
has a closed range and satisfies dim N(T) = oo and dim Y/R(T) <oo. Con-
sider the adjoint operator T*: X* and note that from Theorem 2.13
we have dim N(T*) = dim {Y/R(T)]* < oo. Moreover, since R(T) is closed,
the identity X*/R(T*) = [N(T)J* holds; see Exercise 12 in Section 2.1.
From this identit3r it follows that dim X*/R(T*) 00. By the preceding
case, there exists some E> 0 such that every operator A E (Y*, X*) with
All has a closed range, dim N(A) <oo and dim X*/R(A) = Do.
Now let an operator S e (X, Y) satisfy liT < Then we have
LIT* < This implies that S* has a closed range (and hence S
likewise has a closed range), dim N(S*) < Do and dim X*/R(S*) =
From dim [Y/R(S)]* = dimN(S*) < Do, we infer that dimY/R(S) < Do.
Also, from X*/R(S*) = [N(S)]*, it follows that dim N(S) = Do. The proof
is finished.
i: Fred(X,Y) ZU{oo,oo}
is continuous. Moreover, Fred(X,Y) is an open subset of Fred(X,Y).
Proof. (1) (2)In this case, the closed vector subspaces V = (Ix P)(X)
and TV = (ly Q) (Y) satisf3r the desired properties.
4.4. Fredholm Operators 167
(3) Assume that TST = T for some operator S e (Y, X). Then
(1)
(ST)2 = S(TST) = ST, and similarly (TS)2 = TS. That is, the operators
Pi = ST e (X) and Q = TS e (Y) are projections on X and Y,
respectively. We claim that N(P1) = N(T) and R(Q) = R(T).
For the first identity, note that N(T) c N(ST) and that x e N(ST)
implies Tx = T(STx) = 0, i.e., N(ST) N(T) is also true. For the second
identity, observe that the inclusion R(TS) R(T) is obvious and that for
each x E X we have Tx = TS(Tx) e R(TS). If we let P = P1, then P
is a bounded projection on X and satisfies P(X) = R(P) = N(T).
Proof. The first part follows immediately from Lemma 4.52. For the second
part, let S: Y + X be a pseudoinverse of T, Le., TST = T. Then, by part
(2) of Theorem 4.43, the operator ST is Fredholm. But then, by the same
argument, S is a operator. Now part (1) of Theorem 4.43 yields
i(T) = i(TST)= i(T) + i(S) + i(T). This implies i(S) = i(T).
Lemma 4.55. For any Fredholm operator T E (X) we have the following.
(1) The vector subspace Y = R(T72) is a T-invariant closed sub.
space of X satisfying T(Y) = Y.
(2) If A 0, then N(A T) = N((A T)Jy) = N(A Tjy).
(3) There exists some > 0 such that:
(a) For each A T is Fredhoim.
(b) The nullity n(A T) and the defect d(A T) are constant on
the set {A e C: 0 < AJ <}.
Exercises
L Prove Lemma 4.40. [HINT: If T: X + Y is a bounded operator between
two real Banach spaces, then note that = =
and R(T) =
2. Give a direct proof of Lemma 4.44 without using the HahnBanach
theorem. Also, present an example of an infinite dimensional proper
closed subspace Y of a Banach space X such that for each unit vector
x E X \ Y we have d(x, Y) < 1. [HINT: FLx some x0 E X \ Y and let
d = d(xo, Y) > 0. Next, select some Yo E Y with Mxo poll < and let
= Clearly, I and for any y E Y we have
> i E.
d
x
y IIxo-yoII >
IIxo-voII IIXo-yolI
Proof. To establish this result we need the following simple property whose
proof is left as an exercise.
(P) If X is an infinite dimensional vector subspace of 4, then for each
m the vector subspace
Xm={xEX:
X is also infinite dimensional.
We denote by the standard projection of onto the first n coordinates
of 4 That is, = (yi,. .. , ye We also let
= + -
-
>
172 4. Special Classes of Operators
= +
nu/pTh
MTxII = =
= MS(1x2 = =
i=n+1 i=n-f-1
in+1 i=n+1
Therefore, and so S
S= is a compact operator, and the proof of the first part is finished.
(2) Assume that T is a strictly singular operator. Then the restriction
of T to any infinite dimensional subspace of X is not an isomorphism, and
under this condition all arguments of the above proof remain valid for each
closed infinite dimensional subspace X1 of X.
174 - 4. Special Classes of Operators
Proof. The implication (1) (2) follows immediately from part (2) of
Lemma 4.59. The implication (2) (3) is obvious.
Exercises
1. Show that a bounded operator T: X Y between Banach spaces is
strictly singular if and only if T is not bounded below on any infinite
dimensional closed subspace of X.
4.5. Strictly Singular Operators 177
Integral Operators
179
180 5. Integral Operators
Tx(t)= I (*)
Js
holds for v-almost all t E T.
y(t)=
JS
is the same for any x1 E [x]. This not only explains why T[x] = [y] but it
also shows that we can identify the class T[x] with the function y and write
y = Tx. In sum, an integral operator T: X Y gives rise automatically to
an operator from equivalence classes to functions, i.e., it is a "lifting" or a
"selector."
It should be emphasized that the v-null set of points outside of which the
equality (*) holds depends upon the function x; see Exercise 6 at
the section. The x v-measurable function T(.,.): S x T is referred to
as the kernel of the integral operator T.' It is a common practice to denote
the kernel of an integral operator T by the same letter T, and we will follow
this practice. This is very efficient and when the reader gets accustomed to
this convention, it will cause no problem. We shall also say that a x v-
measurable function T(.,.): S x T defines an integral operator
from X toY if for each x E X the function T(.,t)x(.) belongs to L1(1i) for
v-almost all t E T and the function Tx as given by (*) in Definition 5.1
belongs to Y.
We shall also adhere to the following agreement: An arbitrary point of
S will be denoted exclusively by the letter s and an arbitrary point of T
by the letter t. Accordingly, f(s) indicates a function f: S TR and g(t)
such that for each x E E the function s ' K(s, t)x(s) belongs to L1 (Si,
for v-almost all t T and the equality
Tx(t)= /
Js1
holds for v-almost all C E T1. In other words, when we deal with integral
operators, we can replace S by and T by 71 and assume without loss of
generality that:
For this reason, the integral operators are also known as kernel operators
2As we shall see in Theorem this property characterizes the integral operators.
182 5. Integral Operators
Tx(t)
= f
$
T(s, t)x(s)
f
= $\Dt
T(s, t)x(s) 0.
This proves that T is a positive operator. Clearly, this proof remains valid
for arbitrary subspaces E and F of Lo(,a) and Lo(v), respectivelyand also
for arbitrary measures and ii.
For the converse, assume that T is a positive operator. We must show
that T(s,t) 0 for x v-almost all (s,t) E S x Since S and T are
both o-finite measure spaces, we can assume without loss of generality that
both measures are Moreover, since E and F are order dense ideals in
and Lo(v), respectively, we can also assume in view of Corollary 1.91
that Xs E E and XT E F. Our last simplifying assumption is that the
kernel can be assumed to be x v-integrable. To see this, note
that since the function s F-4 T(s,t) = belongs to L1(,a) for
v-almost all t E T, by deleting a v-null set from T we can suppose that
J'5 T(s, t)) d,a(s) <oo for all t T. So, we can write T = Bk, where
Bk = {t E T: JT(s, d,a(s) k}. Since the v-measurable function
g(t) T(s, t)I d,a(s) is bounded on Bk, we can conclude from Tonelli's
theorem that T(.,.) is integrable over S x Bk for each k. Thus, we can
assume that the function T(.,.) is x v-integrable over S x T.
fAxB
for each A E and B E Since XA E E and T is a positive operator,
we have TXA 0 and consequently fB TXA(t) dv(t) 0. By Definition 5.1
the equality
TXA(t) =
f x v)
= L [f T(s, t)
dv(t) L TXA(t) dv(t) 0,
as desired.
x B)
= f x v)
= L [f T(s, t) dv(t)
= f TXA(t)dv(t).
JB
A direct verification shows that 7rT is a measure on the semiring F.
Next, using the operator 8, we define a new set function 7rs: F * [0, oo)
via the formula
irs(A x B) L SXA(t) dv(t) dv(t).
= IT XB(t)SXA(t)
Clearly, 0 We claim that '/rs is a measure. To verify this, it
suffices to show (since irs is dominated by the measure lrT) that is finitely
additive; see Exercise 15 in Section 1.6.
5.1. The Basics of Integral Operators 185
i=1
B)
= f XB(t) (SXA)(t) =
IT
XB1 (t)(SXA1) (t) du(t)
= >'/rs(AixBi).
Therefore, IrS is indeed finitely additiveand hence it is a measure. A
straightforward verification shows that the outer measures and satisfy
0 and that, when restricted to 0 they are both cr-finite.
Next, observe that the measure restricted to the cr-algebra
is absolutely continuous with respect to the cr-finite measure x ii. From
it follows that is also absolutely continuous with respect to
x ii. But then, by the classical theorem, there exists
a function S(.,.): S x T > [0, oo) (and hence
is also x ui-measurable) such that = S x ii) holds for each
x B)
= f (SXA)(t) du(t)
= f S(s,t) x v)
= du(t).
Now fix some A E with (A) <oo and consider the collection
A = {B E is integrable over A x B}.
Clearly,A is a semiring such that the cr-algebra generated by this semiring
coincides with see the proof of Theorem From this and Lemma 1.96,
186 5. Integral Operators
SXA(t)
To finish the proof, note first that the operator S is order continuous.
(This follows from the inequality 0 S T and the fact that T
order is
continuous by Corollary 5.4.) This and (**) imply immediately that for each
x E E we have Sx(t) = f8 S(s, t)x(s) d,ti(s) for all t E T. That is,
for
[TIx(t)
Proof. Let T: E
F be a regular integral operator. Then, by Theo-
rem 5.11, its modulus TI: E F is also an integral operator, and so by
Corollary 5.4 the operator T itself
is also order continuous.
to this, the regular integral operators are also known as absolute integral operators
188 - 5. Integral Operators
The last statement of the corollary follows from the previous part since
each integral operator T: E F considered as an operator from E to L0 (ii)
is regular.
The kernel T(., on S x T is defined as follows. For each s e [0, 1] and for
each t e where m = 1, 2,..., we let
T(s, t) sin(2'irns).
It is obvious that T(.,.) is a bounded x v-measurable function on S x T.
First, we verify that the integral operator T generated by this kernel
maps E to F. To see this, note that for each x E E and each t e we
have Tx(t) = f5 x(s) sin(27rns) ds, and so the RiemannLebesgue lemma
yields
dt x(s) ds 0
IT =
Thus, y = Tx e F.
Next, we claim that the integral operator TI with kernel T(., .)J does not
map E to F. To this end, it suffices to show that the function v =
does not belong to F. Indeed, for each t e we have
v(t)
= f sin(27rns) ds f sin2 (2irns) ds
Proof. We can assume without loss of generality that E and F are order
dense in Lo(1i) and Lo(v), respectively. Theorems 5.9 and 5.11 guarantee
that K(E,F) is an ideal in 4(E,F) and that the above formulas for cal-
culating S V T and S A T are true. The only thing left to be shown is that
K(E,F) is order closed in 4(E,F).
To see this, assume that a net {Ta} of positive integral operators sat-
isfies Ta I S in 4(E,F). By Corollary 1.91, we can assume without loss
of generality that is E E, and so Ta15 I S's in F. Since F has the
countable sup property, there is an increasing sequence of indices
such that Taml5 I Sls, that is, f5Tam(5,t)dIl(S) I Sls(t) for v-almost
all t E T. By Theorem 5.5, the kernels Tan (s, t) satisfy Tan (s, t) I. This
implies that there exists a function T(.,.) such that Tan (s, t) I T(s, t) for
x v-almost all (s, t) E S x T. Without loss of generality, we can suppose
that Tan(S,t) I T(s,t) for each (s,t) E S x T. Next, we claim that T(s,t)
is finite for x all (s, t) E S x T.
Sx(t) = lim
12-400
= fl-400
lim f Tam(S, t)x(s) d/2(s)
= f T(s, t)x(s) d/2(s).
5. Integral Operators
So, by B. Levi's theorem, f8To(s,t)x(s)dji(s) Tx(t) < oo, that is, the
function s To(s, t)x(s) belongs to L1 (ji) for ti-almost all t E as required.
5.1. The Basics of Integral Operators 191
Furthermore, using once more that To(s, I To(s, t)x(s), we see that
f To(s,t)x(s)
The latter and (*) yield Tx(t) = f5 To(s, t)x(s) each t B. That for
is indeed a kernel operator with the same kernel as T0.
is, T
To show that the assumption that T is a positive operator is not essential,
we proceed as follows. Since T: E Lo(zi) is u-order continuous, T is
regular (see Corollary 1.89(a)), and hence the operator T+: E Lo(zi)
exists and is u-order continuous. By Theorem 5.10, the integral operator
T0: E0 Lo(z') is also an integral operator
with kernel .). It remains to note that the restriction of to E0
coincides with Thus, instead ofT we can consider the positive operators
T+ and T,
and the desired conclusion follows.
Exercises
1. Assume that ,a and ii are a-finite measures. Show in detail that if the
kernels of two integral operators 8, T: E F satisfy 8(s, t) =T(s, t) for
x v-almost all (s, t) c S x T, then 8x = Tx for each x C E.
2. Let (5, ,a) and (T, ii) be two a-finite measure spaces, and consider
a ,a x v-measurable function K: S x T p R. Assume that a function
f C Lo(ji) satisfies K(.,t)f(.) C L1(ji) for v-almost all t C T. Show that
the function g: T R defined v-almost everywhere by the formula
g(t)
= f K(s, t)f(s) dy(s)
and let 1 <p, q < satisfy +=1. Show that an integral operator
T: maps the closed unit ball of to an order bounded
subset of Lo(v) (i.e., there exists a function g E Lo(v) such that g
for each s e with 1) if and only if there exists a v-null set
BE such that
f
for each t E T\B.
In particular, an integral operator T: L2 (ii) + (ii) is a Carleman
operator if and only if T maps the closed unit ball of L2(1i) into an order
bounded subset of Lo(v).
6. Give an example of an integral operator on L2 [0, 1J which is not a Car-
lernan operatorwhere [0, 1] is equipped with its Lebesgue measure. In
view of Exercise 4, this demonstrates that in general the exceptional set in
the definition of an integral operator depends on the function the operator
is acting upon. [ifiNT: Use Exercise 3 above.]
7. Let be an increasing sequence of regular integral operators and let T
be another regular integral operator. Show that T holds in F)
if and only if T(s, t) t) x CF.
8. Let E be an ideal in Lo(,u), where (8, is a non-atomic cr-finite mea-
sure space. If T: E + E is a regular integral operator, then show that T
is disjoint from the identity operator, i.e., ITt A I = 0.
9. (AbramovichAliprantisZame [20]) Show that for a Dedekind complete
AM-space E with unit e the following two statements are equivalent.
(a) There exists a probability measure space and a surjective
lattice isometry T: E + with Te 1.
(b) E admits a strictly positive order continuous linear functional.
[HINT: Assume that E + R is a strictly positive order continuous
linear functional. We can assume that gb(e) 1. Clearly, the formula
Us HI = defines an L-norm on E. If E is the norm completion of
(E, I
then E is an AL-space (see Exercise 6 of Section 3.4) having
e as a weak unit. Now, according to the KakutaniBohnenblustNalcano
Representation Theorem 3.5, E is lattice isometric to some L1 (7r)-space,
where e corresponds to the constant function 1. Show that is a proba-
bility measure and that E is lattice isometric to I
10. Let (8, ,u) and (T, v) be two arbitrary measure spaces, and assume
that 1 <p, q < cc satisfy + = 1. If a function T(.,.) e x
then show that the formula
Ts(t) t)s(s) dy(s)
= f T(s,
5.2. Abstract Integral Operators 193
Proof. Clearly, we can assume without loss of generality that the ideal G
is order dense in Lo(7r). By Corollary 1.91 there exists a disjoint sequence
C E such that = and the function belongs to C
for each n E N. Then the series u = >t1 where
denotes the norm of C, is norm convergent in C. By Lemma 5.21, we have
u(w) = (w) for 7r-almost all w E 11. This implies that
II)
u is a weak unit. D
Theorem 5.23. Every integral operator between Banach function spaces is
norm continuous.
Our next goal is to establish the important fact (due to G. Ya. Lozanov-
sky) that each normed function space is necessarily normal. To prove this,
we need the following technical result due to W. A. J. Luxemburg and
A. C. Zaanen [224, Lemma 31.1].
196 5. Integral Operators
Proof. By hypothesis 0 u holds for some u E L and for each ri. Fix
0, and for each n and m let
un,m =
u and the positivity of b, we see that cb(un,m) Tm b(u).
So, there is a strictly increasing sequence of natural numbers such
that
0 cb(un,m) <* (*)
for all m Next, for each ri let
Zn = ul,m1 A 'LL2,rn2 A
b(yn) = <.
claim that un
We + Yn for each n. To establish this inequality, it
suffices to prove that un Uj,rrij + Yn for each j = 1, . , n. For this, note
.
+ Yn = + Ui,mj)
+ Uj,mj) =
Finally, using the Riesz Decomposition Property, we can write each
as un = +Wn, where 0 Vn Zn and 0 Wn Yri. Clearly, the sequences
and {Wn} satisfy the desired properties.
Theorem 525 (Lozanovsky). If C is a normed function space, then
separates the points of C. In particular, C is a normal RiesZ space.
5.2. Abstract Integral Operators 197
yEH and Oy
where denotes the lattice norm of C.
We claim that p(x) > 0 for each 0 < x E H. To see this, assume by way
of contradiction that p(x) = 0 for some 0 < x E H. Consequently, there
exists a sequence [0, x] satisfying + b(x < for each
n. From Jun11 < and Lemma 5.24 (applied to it follows that
given any > 0 there exist two sequences and such that 0,
<e and = + for each n. Then for each n we have
=
<
The function
= f
is uniquely determined on the carrier of C.
If C is an order dense normal function space in LO(7r), then the func-
tion is unique and the mapping from to Lo(7r), is a lattice
isomorphism whose image C' is the order dense normal function space in
LO(7r) described by
Moreover, C' can be identified with where each function g E C' acts on
C as an order continuous linear functional via the formula
thisreason the order continuous linear functionals are also called normal integrals.
6 ideal is often referred to as the Kthe dual of C.
5.2. Abstract Integral Operators 199
This
= lim
f
= lim
establishes the representation of
= f
The verification of the uniqueness
of on the carrier of C is straightforward.
Assume now that C is normal and order dense. The uniqueness of gi,
follows easily from the fact that in this case the carrier of C is all of
Moreover, it should be clear that the mapping 'k is linear and that its
image C' is an ideal in L0(ir). To see that 'k is a lattice isomorphism,
fix E Then for each x E we have
= sup = sup
f = f
This implies that = and so 'k is a lattice isomorphism.
-For the last part, we need to verify that is order dense in Lo(ir).
To this end, take any 0 < y E L0(ir), and then choose some x E E+
satisfying 0 < x y. Since the order continuous dual separates the
points of C, there exists some 0 <b E (with the corresponding function
o < E Lo(7r)) such that b(x) = > 0. This implies
o Ax x y, and from Ax E it follows that is order dense
in L0(ir). The proof is finished.
UCUAnXBn. (t)
Now recall that the order denseness of E implies (according to Theorem 5.26)
that is also order dense. In particular, it follows from Corollary 1.91 that
for each m there exist sets As',... with E and
v-measurable sets Bk,... with e F such that =
and = This shows that in the union in (f), we can suppose that
E and e F for each n. Next, put = Ufl Ak x Bk) and
note that the integral operator with kernel satisfies the inequalities
o xBk Since XAk E 13 for each k, it follows that
e 13. From I Xu we get T T E 13; see Exercise 7 in
Section 5.1.
Finally, let T: E F be a positive integral operator with kernel T(.,.).
In view of Theorem 5.5, we can assume without loss of generality that
T(s, t) > 0 for all (s, t) E S x T. Since the measure x ii is o-finite, there
ists a sequence of x v-step functions such that 0 t) I T(s, t)
for each (s, t) E S x T. Now notice that if = is a x
function satisfying 0 T, then each x v-measurable set satis-
fies the conditions of the previous part. Consequently, the integral operator
generated by belongs to 13, and so the integral operator with kernel
also belongs to 13. Thus, the integral operator with kernel 'J!Th belongs
to 13. Since 13 is a band, it follows (again by Exercise 7 in Section 5.1) that
T e 13. Therefore, F) 13 is also true, and the proof is finished.
Recall that if T: L + M is an order continuous regular operator be-
tween two Riesz spaces with M Dedekind complete, then its order adjoint
operator M" 17, defined for each x E L and each y e via the
duality identity
(Tx, y) = (x, =
carries order continuous linear functionals to order continuous linear func-
tionals, i.e., holds. To see this, let y E and assume
Xcy 0 holds in L. Then the order continuity of T implies TXa 0 in M,
________-
(Tx, y)
= L [f T(s, t)x(s) y(t) dv(t)
dv(t).
So, by Tonelli's Theorem 1.98, the function (s, t) T(s, t)x(s)y(t) is x ii-
integrable. In particular, for each fixed x E the function T(s, .)x(s)y(.)
is integrable over T for 'u-almost all s E 5, and the function
g(s) = LT(5,t)Y(t)x(5)dt'(t) =
(Tx,y) =
= liminf
<
lim sup
ntoo
urn sup
fl4 00
= and CCA}.
Notice that for each measurable subset C of A we have
SXC+ii(A\C)eSXA+ii(A)e [1+ii(A)]e. (t)
F has the countable sup property, there is a sequence
Since
such that C A for each n and
inf + \ E and C A} =0. (*)
inf
mEN
+ ii(A \ )e} =0.
If this claim is established, then we get a contradiction as follows. Since
\ Cam) 0,it follows from 3(b) that SXA\Cnm 0 or, equivalently,
inf
flENk
and cA} =0.
In particular, we have that infflENk = 0, infflENk \ = 0, and
that each set Nki5 infinite. Let Nk = {mk,1, mk12,. .}, where mk,n
.
for each n.
Now for each k and r in N put
Yk,r
=
From = it follows that for each k we have Yk,r
0, 0. Conse-
quently, by the diagonal property (see Theorem 1.87) there exists a sequence
{Yk} in Lo(v) with Yk t 0 and a strictly increasing sequence {rk} of natural
numbers such that Yk,rk Yk for each k. Let It'I = {mk,1,. .. ,
and note that M is a countable subset on N. Therefore, M = {ni, n2,. .}, .
To see this, fix 6> 0, and then pick some mU such that p..(A \ Cam) <6 for
each m mU. Then for each k we have:
g < inf
mmo
+ \ Cnm)e} inf + 6e Yk + 6e,
where the last inequality follows from the definition of functions Yk,r and the
fact that Yk,rk But then Yk 0 yields g <6e. Since 6 > 0 is arbitrary,
we get g = 0. This completes the proof of the theorem.
function spaces with trivial order dual may carry only trivial integral opera-
tors. For instance, it follows from a well known theorem by S. Kwapien [1963
(see also [170, Theorem 8.4]) that the function space Lo[0, 1] admits only
the trivial integral operator.
The first simple application of Theorem 5.30 describes an important
class of Banach function spaces on which every continuous operator is an
integral operator.
Tx(t) = v(t)
f K(s, t)u(s)x(s)
for v-almost all t e T.
isI
Tx(t) = v(t) K(s,t)u(s)x(s) dy(s)
for v-almost all t T.
Exercises
1. Let E and C be two Banach function spaces associated with Lo(7r). Show
that E fl C is also a Banach function spacewhich is order dense if both
E and F are order dense ideals. [HINT: If IE and denote the
1
Show that (a) (b), (c) (d), and (b) (c). Does (d) (b)?
[HINT: Use Exercise 4 in Section 5.1 and Lemma 1.30(c).]
9. If is an order bounded positive sequence in some Banach function
space E in L0(71), then prove that 0 is equivalent to 0,
i.e., to the property: Xn(W)y(W) d7r(w) 0 for each y E
Is this equivalence true if is not assumed to be a positive se-
quence?
10. If E and F are Banach function spaces, then show that is a
principal band in Lr(E, F). [HINT: We can assume that E and F are
order dense ideals. If e and cb are weak units in F and respectively,
then show that =
Prove Corollary 5.31 using Theorem 5.28 instead of Theorem 5.30.
Show that Dunford's Theorem (see Corollary 5.33) is not valid for continu-
ous operators between arbitrary Li-spaces. [HINT: The identity operator
I: L1 [0, 1] L1 [0, 1] is disjoint from every regular integral operator; see
Exercise 8 in Section 5.1.]
13. Assume that (8, ,u), (T, ii), and (Q, 0) are three u-finite mea-
sure spaces and that E, F, and G are ideals in L0(1i), Lo(v), and L0(71),
respective'y. Assume also that in the scheme E F C the operators
S and T are both regular and order continuous.
If either S or T is an integral operator, then show that TS is likewise
an integral operator.
14. Assume that (8, ,u) and (T, ii) are two u-finite measure spaces
with (8, ,ii) non-atomic. Also, suppose that E and F are ideals in
L0(8, and L0(T, ii), respectively, and letS: E F be a regular
integral operator.
210 5. Integral Operators
15. Assume that (8, (T, ii), and (Q, 8) are three a-finite mea
sure spaces and that E, F, and G are ideals in L0 (/2), Lo (ii), and L0 (8),
respectively. Suppose also that in the scheme E F G both S and T
are regular integral operators with kernels and T(., '), respectively.
Show that the composition operator R = TS is also an integral
ator whose kernel is given by the convolution of the two kernels S(.,.)
and T(., .), that is,
I
lIp-norm of f in The same conclusion is true for p cc. That is,
Ill = for each function I Thus, for each 1 <p oo
the Banach lattice (A) is a closed vector sublattice of (>) containing
the constant functions. We shall see later in this section that every closed
vector sublattice of an that includes the function 1 is of this
type.
Before proceeding further, let us make one more cautionary comment
about the Banach lattice where A is a cr-subalgebra of >1 If we
consider the probability measure space A, p) in its own right, then we
2 12 5. Integral Operators
can think of (A) as the usual associated with the measure space
( A, ,u). On the other hand, if we consider as a vector sublattice
of then we view the classes of as classes in That is,
we tacitly identify (A) with its image in under the lattice isometry
[hA The following example clarifies the situation. Let
. = [0, 1],
let be the of all Lebesgue measurable subsets of [0, 1], and let
,u be the Lebesgue measure. Consider also the a-subalgebra of given by
A = {, [0, 1], [0, 1] }. Then it is easy to see that
= {clX[OI} + C2 E
7RecalI (see Theorem 1 80 and the paragraph thereafter) that since we deal here with a finite
measure space, each ideal E in Lo(E) satisfies the countable sup Accordingly, a positive
operator T: E + E is order continuous if and only if 0 implies 0.
5.3. Conditional Expectations and Positive Projections 213
o (t)
Since fP E L1(>J), the function belongs to Li(A) There-
fore, from (t) we get E L1(>J), that is, E(f[A) E This
proves that E(.[A) leaves invariant.
=
Therefore, E(. A): is a contraction, as claimed.
Following the work of G. Birkhoff [63], we shall say that a bounded
operator T: L1(>J) Li(>J) satisfies the averaging property (or that
the operator T is an averaging operator) if for each pair h, f
such that hTf E L1 [in particular, this is so whenever h E (>J) and
f we have
T(hTf) =ThTf.
Conditional expectations enjoy the averaging property.
8 Positive projections leaving the function 1 fixed are sometimes called Markov projections.
This is in agreement with our Definition 3.15 of a Markov operator
214 5. Integral Operators
I
JA JA
for each A E A or, equivalently,
/
JAnD
XB = f
JAnD
E(xB IA)
Proof. The implications (1) (2) (3) are trivial. Hence, only the
implication (3) (1) needs a proof. So, assume that (3) is true. Consider
the collection of sets
A={AETh XA"}
5.3. Conditional Expectations and Positive Projections 215
XACIXA,
we conclude at once that A is a subalgebra of E. To see that A is a a-algebra,
assume that some sequence A satisfies I A. This certainly
implies that A E and XA E E, our condition (3)
guarantees that XA E Y. Thus, A E A, and so A is a a-subalgebra of
Since the vector sublattice of all A-step functions is included in Y and
is order dense in E fl L0 (A), it follows from (3) that E fl L0(A) C Y. Now
let f Y and consider the set A {w E f(w) > Since Y is a
vector sublattice and 1 E Y, it follows that A 1} Y and moreover
mf+ A I I Using once more (3), we see that Y and so A E A.
Finally, note that since I ci E Y for each scalar c, the preceding case
implies
Proof. It should be clear that (1) implies (2). Now assume that (2) is
true. Since norm convergence in implies order convergence, it easily
follows that Y is a normed closed subalgebra of But then it is well
known from the proof of the StoneWeierstrass theorem (see for instance [31,
Theorem 11.5, p. 89]) that Y is a vector sublattice. The validity of (1) now
follows from Lemma 5.40. 0
The next result presents two basic order properties of contractive oper-
ators on
216 5. Integral Operators
= = = 1.
This guarantees the existence of some constant c> 0 such that = ci;
see for instance [32, Problem 31.4]. Clearly, c = 1, and from this and (**)
we get T*1 = 1.
(b)Consider first the case p = oo. Pick any 0 x e Dividing
by if necessary, we can assume that 0 x $ 1, and 50 0 1 x 1.
Taking into account that = 1 and that the closed unit ball of
is the order interval [1, 1], it follows that 1 Tx = T(1 x) e [1, 1].
implies that the second adjoint T** of T is also a positive operator. Since
T** agrees with T on L1 (E), it follows that T itself is positive too.
Exercise 7 at the end of the section shows that for other values of p
the conclusions of the preceding lemma do not hold in general. Recall that
a Banach lattice E is said to have strictly monotone norm if xl <
implies lix 11 < IIylI. The Lu-spaces with 1 p < 00 have strictly monotone
norms.
Lemma 5.43. Let T: E * E be an operator on a vector lattice and assume
that, either
(a) T is a strictly positive projection, or
(b) E is a Banach lattice with strictly monotone norm and T is a con-
tractive positive operator.
Then the fixed space 2(T) = {x E: Tx = x} of the operator T is a
e
vector sublattice of E (which, in case (b), is also norm closed).
Conditional Expectations' and Positive Projections 217
Proof. (a) To begin with, notice that since T is a projection, the fixed space
of T simply coincides with the range T(E) of T.
Now take any x E T(E). Then lxi TIxj, and so TIxI lxi 0.
Since T2 = T, it follows that jxj) = 0, and so the strict positivity
of T implies xl, that is, lxi E Therefore, = T(E) is a
vector sublattice of E.
IA = f = =
f P(xxj (*)
Conditional Expectations and Positive Projections 219
Using that P satisfies the averaging property and that = XA7 we get
P(xxA) = P(XPXA) = P(x)P(xA) = P(x)XA
for each x E E. From this and (*), we obtain
f
JA J
[ JA
Since Px is and for each x E E, it follows from
the definition of the conditional expectation that
E E. In particular, E(q5[A) = = P1 = 1. This establishes the
desired representation of P in the special case when E
CASE II: E is an ideal such that (E) E L1 (E).
The hypotheses that P is positive and P1 = 1 guarantee that P leaves
invariant and = 1. So, if denotes the restriction of P to
then is a contractive projection on satisfying = 1.
We claim that the range of is a vector sublattice of (s). To see
this, let x, y and put z PxVPy E Since the range of the
operator P: E E is a vector sublattice of E, there exists some u E E such
that Pu = z. But then the function z E satisfies z Pz
This shows that the range of is a vector sublattice of
According to Case I, there exist a a-subalgebra A of and a non-
negative function q5 E such that = E
This means that the desired representation
Px E(bxjA)
is valid for each x E Since is an order dense ideal in L1(>), it
follows that is also an order dense ideal in E. Now use the fact that
P and are both order continuous on E to infer that Px =
E E.
The obtained representation implies, in particular, that for each x E E
the function q5x is integrable. This means that q5 E Eq'.
(2) Assume that there exist a a-subalgebra A of and a non-
(1)
negative function q5 such that = 1 and Px
E E. In particular, notice that = q51 E Lemma 5.37
guarantees that P is a positive order continuous operator. We also have
P1 = = = 1 by the condition on
Let us verify that P is a projectiom Keeping in mind that E(' IA) satisfies
the averaging property, for each x E E we obtain
P2x =
= = = Px.
220 5. Integral Operators
Proof. Only the implication (1) (2) needs a proof. Since P is strictly
positive, it follows from Lemma 5.43(a) that the range Y = P(E) is a
vector sublattice of E. Therefore, Theorem 5.45 applies and guarantees the
existence of a a non-negative function q5 E such
that S(q5JA) = 1 and Px = S(q5xIA) for each x E E. To see that b is strictly
positive, let A e E with > 0. Then 0 < XA E E, and by the strict
positivity of P we get > 0. Therefore,
IA =
holds for all A E E with
f = f [A)
= f PXA >0
f
For the converse, assume that f x d1i = Tx d1i holds for each x E E.
Rewriting this identity in duality form and taking into account that I e
we see that
(x,1) = (Tx,1) =
(x, I I) = 0 for each x E E. This implies I = I.
(2) We show first that T is strictly positive. To this end, let 0 < x E E.
From part (1), we have f Tx = f x dpi> 0. This implies Tx > 0, and so
T is strictly positive.
For the order continuity of T, assume 0 in E. This implies 0
in L1(E), and so from part (1) we get
/
Jci
t,. 0 ,u-a.e. Consequently,
T is order continuous.
We now present the definition of a bistochastic projection.
Definition 5.49. Let E be an ideal in L1 (E) containing the constants. A
positive projection P: E B is said to be bistochastic if both P and its
order adjoint operator are Markov, that is, P1 = I and
=
which is a contradiction. Therefore, q5 1 and so P = e(.IA).
If P is a contractive projection on L1(E) satisfying P1 = then by 1,
Lemma 5.42(a) the projection P is bistochastic, and so the preceding corol-
lary immediately yields the classical result of R. G. Douglas [112].
Corollary 5.52 (Douglas). For an operator P: Li(E) L1(E) the follow.-
ing statements are equivalent.
(1) P is a contractive projection satisfying P1 = 1.
(2) P is a conditional expectation operator. That is, there exists a
unique A of E such that P = e(.jA).
Corollary 5.52 was generalized to for I <p < oo with p 2
by T. And Under the assumption that the contractive projection is
positive the restriction p 2 is redundant. We present below a very simple
proof of this result.
Corollary 5.53 (And). For a positive operator P: where
I <p < 00, the following statements are equivalent.
(1) P is a contractive projection satisfying P1 = 1.
Proof. Notice that for p = 1 the result is simply Douglas' result (Corol-
lary 5.52). So, we can assume that 1 < p < oo.
(
1) The projection P is certainly order continuous since the
(2)
norm on is order continuous. In view of Lemma 5.43(b), the range
of P is a vector sublattice of since the operator P is contractive
and the norm on is strictly monotone. Hence, by Theorem 545, the
representation Px = holds. We want to show that = 1. For
this to be true it suffices to know that P*1 = 1. But this is guaranteed by
Lemma
(2) (1) The implication was established in Lemma 5.38.
seems that C. Moy [2461 was the first who used the conditional expec-
tations for representations of some operators that were not assumed to be
projections. Some considerable generalizations of his results were obtained
by R G. Dodds, C. B. Ruijsmans and B. de Pagter [110]; see for instance
Exercise 10 at the end of the section. For a current survey on contractive
projections-on Banach spaces see the article by B. Randrianantoanina [274].
Our next comment relates some of the results obtained in this sections
with the interpolation of operators on Banach function spaces. For this
comment we need some terminology from the theory of Banach function
spaces.
o Two functions x, y E Lo(,u) are said to be equirneasurable if they
generate the same distribution function, if the equality
ft x(w) <t}) = E ft y(w) t})
for each t R.
holds
o A Banach function space E associated with some is
called rearrangement invariant if every y E L0 that is
to some x E E belongs to E and = lxii.
224 5. Integral Operators
Px =
Our next example shows that in general a 1\'Iarkov projection can easily
fail to be contractive and its range can fail to be a vector sublattice.
Example 5.56. We shall establish the existence of a positive projection
P: L1[O, 1] L1[O, 1] such that:
(1) P1=1.
(2) >1.
(3) The range of P is not a vector sub1attice of L1 [0, 1].
= {c1XA +
Li(A) C2XB + c3 E R },
and the positive operator Li(>2) is given by
and has L1 (A) as its range. (See Exercise 4 at the end of the section.)
Next, define Q: Li(A) Li(A) by
We conclude the section with one more important property of the con-
ditional expectation operator asserting that the operator S('IA) is basically
seif-adjoint.
Theorem 5.57. For each a-subalgebra A of the cr-algebra the adjoint
operator S(.IA)*: satisfies
E That is, coincides with the restriction of the conditional
ezpectation operator S('IA) to the space
226 5. Integral Operators
is enough to prove that and have the same range and the
same kernel. Since (by Lemma 5.38) both projections and
are contractions on it follows that E(.jA) and are orthogonal
projections; see Exercise 14 at the end of the section. In particular, if
E('IA) and have the same range, then they automatically have the
same kernelwhich is the orthogonal complement of the common range.
Consequently, E(' A) = E(. if and only if E (. IA) and E have the same
range as projections on To verify the latter, it suffices to establish
that XA E L2(B) for each A e A and XB E L2(A) for each B E 13.
Let A e A. Using the definition of the adjoint operator and the fact
that PXA XA, we have:
(E (XA 13), XA) = XA, XA) = (XA, PXA) = (XA, XA) = (A).
That is, we have obtained the equality in the Holder inequality, and therefore
the function E(xA 13) must be equal to 1 on the set A. Similarly, we can show
that E(XAC 13) must be equal to 1 on the set Since both functions E(XA jB)
and E(XAC 13) are non-negative and E(XA [B) + E(XAc = E(1113) = 1, we
can conclude that XA = e(XA 13) e L2(13).
Exercises
1. Let E, p) be a probability space, and assume that A is a usubalgebra
of E. Show that for each function f we have = 1ff
2. This exercise shows that every Banach lattice with order continuous norm
and a weak unit can be represented by a Banach function space that
"lives" between and L1.
Let E be a Banach lattice with order continuous norm and a weak
unit e. Show that there exists a probability measure space E, p) and
an ideal F in L1(E) containing the constants, i.e., F L1(p),
and a surjective lattice isomorphism T: E + F such that T(e) = 1.
[HINT: Let E JR be a strictly positive 'inear functional satisfying
cb(e) =1. (Since E has order continuous norm, such a functional always
exists;see [30, Theorem 12.14, p. 183].) Now consider the L-norm . J
I =
(c) If is at most countable, then every u-algebra of subsets of is
generated by a unique at most countable partition of ft
(d) Now suppose that E, p) is a probability measure space and that
{ is an at most countable measurable partition of i.e, be-
sides being a partition of we also have E for each
iEI.
have
E(f IA) =
iEI
5. For an arbitrary convex function JR > JR, any function f
E E establish the following
inequalitiesknown as Jensen's
(a) fcbofdp.
(b) A)) E('iL' o in L1(A).
6. Give an example of an ideal E in L1(E) such that E is a Banach lattice
and contains a non-trivial norm closed order dense ideal in E.
7. Show that the conclusions of Lemma 5.42 do not hold in general for the
remaining values of the indices.
228 5. Integral Operators
considered in its own right is a vector lattice! This special feature possessed
by certain subspaces is formalized in the next definition.
Definition 5.58. A vector subspace Y of an ordered vector space X is said
to be a laUice-subspace if Y equipped with the ordering induced by X is a
vector lattice. Equivalently, Y is a lattice-subspace of X if Y equipped with
the cone Y+ = Y fl is a vector lattice.
If Y is a lattice-subspace of an ordered vector space X, then the lattice
operations in the vector lattice (Y, will be denoted by V and A (or by
Vy and Ay, respectively). That is, if yi, E Y, then the supremum of the
set {y1,y2} in the vector lattice (Y,Y) is denoted by ylVy2 or Yl
Similarly, the infimum of the set {yi, Y2} in the vector lattice (Y, 4) is
denoted by ylAy2 or y Y will be
denoted by
It should be obvious that a vector sublattice of a vector lattice is auto-
matically a lattice-subspace. The converse is not true in general; namely,
the class of lattice-subspaces is considerably larger than the class of vector
sublattices, and so, an arbitrary lattice-subspace need not be a vector sub-
lattice. For instance, if we consider the Riesz space E = C[O, 1] and let Y
be the vector subspace of E consisting of all linear functions, i.e., Y consists
of all functions of the form y(t) = mt + b, then Y is a lattice-subspace of E
that fails to be a vector sublattice.
Theorem 5.59 does not characterize lattice-subspaces. That is, not every
lattice-subspace of a vector lattice is the range of a positive projection. This
is demonstrated by the next example taken from [18].
Example 5.61. Consider the positive functions xl, x2 E C[0, 1] defined by
xi(t) =
1(t_1)2
and x2(t) =
1!_t
2
ifO<t<1
2
1 .
Then foP is a positive linear functional on C[O, iJ. So, by the Riesz Repre-
sentation Theorem, there exists a (unique regular Borel measure
v on [0,1] such that [f o P}(x) f[01] x dv for each x E C[O, 1]. Letting
x = X2, we get 0 = [f o P](x2) = f[01] x2 dv, from which it follows that
Supp v = { }. Finally, letting x = Wi, we obtain
1=[foP](xi)=f =0,
[071]
Proof. The verification of this identity follows easily from Lemma 5.62. We
leave the details as an exercise.
=
Moreover, the operator F: p Y is a surjective interval preserving
lattice homomorphism.
= Ypq 0,
andhence
V A A for each p, and thus
\'j/n
W i=1 //\\ 'W p_I /I\\ q_iYpq'
By symmetry, the reverse inequality is also true, and so the validity of (*)
has been established. In other words, the mapping P is well defined.
Next, observe that if x = = Ypq, then
ID
j //\\p1 W q=iYpq'
This follows easily from Lemma 5.62. We are now ready to establish the
basic properties of the mapping P.
o The mapping P is additive.
To see this, let x = and y Ypq, and note that
P(x + y) = + VL Ypq)
= + Ypq))
/ i
'Wi=1 -
Wj=l//\\j=lXijl Wp=1//\\q=1YPq
= P(x)+P(y).
o The mapping P is homogeneous.
Let x = and fix some scalar If 0, then it should be
clear that P(cEx) = aP(x). If < 0, then note that
= = VT=ia'xij
1 DI
= and =
Now fix some vector x = in Then for each i we have
x Xij, and the positivity of Q yields Q(x) =
for each i. This implies Q(x) V A X&j F(x). Since the vector
x can also be written in the form x = a similar argument
shows that Q(x) F(x). Therefore, Q(x) = P(x) for each x E and
the proof is finished.
Exercises
1. If P is a positive projection on an for some 1 p <oo, then
show that its range is again order isomorphic (up to an equivalent norm)
to an for some measure space (Il1, E1,v).
2. Prove Lemma 5.62.
Spectral Properties
This chapter introduces the spectrum of a bounded operator and studies its
basic properties. We start with a general discussion of the spectrum, of the
resolvent set, and of the spectral radius. Subsequently, we introduce and
study the resolvent function. In particular, we pay special attention to the
resolvent function of a positive operatorS
The various important parts of the spectrum (the point spectrum, the
residual spectrum, the continuous spectrum and the approximate point spec-
trum) are all introduced and some of their interrelationships are obtained.
The chapter culminates with an extensive treatment of the "Functional Cal-
culus" with a special emphasis on the Spectral Mapping Theorem and some
of its applications.
The Spectral Mapping Theorem is used in conjunction with the notion of
a spectral set to construct reducing pairs of subspaces for a given operator.
We also study the isolated points in the spectrum that are poles of the
resolvent function.
237
238 6. Spectral Properties
Theorem 6.3. If T E (X) and > 11Th, then,\ E p(T) and the resolvent
T) is given by the Laurent series
wherethe series converges in the operator norm in the Banach space (X).
Moreover, for each ,\1> 11Th we have
= lim
in-*oo
lim
m=0
in in
= lim I
L
lim I
Similarly, S(,\ T) = I, and so S = T). For the norm inequality
note that
00 00
= = IAI-UTII'
I I(
(1) R(14R(\).
(2) IfS e (X) satisfies ST = TS, then SR(.A,T) for
each E In other words, if T belongs to the commutant of
an operator S, then T) belongs to it too. In particular, T and
R(,\,T) commuteS
(3) R(.A) R(p) =
240 6. Spectral Properties
T) = T)S = S =
=
= =
R(,u)
Theorem 6.6. If T E (X) and )'o E p(T), then the open disk
D (,\o, is included in p(T) and for each E we have
00 00
= =
(*)
and
S . (**)
6.1. The Spectrum of an Operator 241
Proof. Let A Aol Then 'y = A Aol' IR(Ao)i1 <1 and so the
series is norm convergent. Now note that
(A T) A)Th(A0
= [(Ao A) + T)]
(A0
[(Ao (A0
Sim:ilarly, A)Th(A0 T) = I.
establishes
This
the representation for R( A). In view of Theorem 1.72, the function R( is
analytic in p(T).
To establish the inequality (*), observe that
IIR(1\)11 = -
IA0 -
=
= IIR(Ao)1!.
R(A) R(A0) =
Corollary 6.8. For every operator T E (X) the derivative of its resolvent
function is given by
dA
Proof. By Corollary 6.7 the resolvent function is continuous. So, using the
resolvent identity, we see that
dR(A,T)
= urn = urn =
as claimed.
r(T)
R(A, T)
=
244 6. Spectral Prop erties
Proof. Let T e (X). Ife o-(T), then we claim that e cr(TT1). Indeed,
if then there exists some bounded operator 8: X X such
that (,V' TTh)8 = 8(,V' TTh) = I. Write
= +
and note that the operators A + + + + Ta') 8 and
B= satisfy T)A = B(.AT) = I.
This implies )'.. e p(T), which is a contradiction.1
Now choose some E cr(T) with = r(T). From e we see
that and so = r(T) holds for each m. It follows
that
r(T) inf <liminf (*)
Note that the resolvent function R(,\, T) is analytic on the open annulus
A= e C: > r(T)} and that A is the largest open annulus centered
at zero that is included in p(T). By Laurent's Theorem 1.78 we must have
= for each > r(T), and from Theorem 1.77
we see that r(T) = lim A glance at (*) guarantees that
r(T) = =
Werefer to Exercise 6 at the end of this section elementary proof
of the existence of the limit *.
Definition 6.13. An operator T E (X) is said to be:
(1) nilpotent if = 0 for some positive integer k and
(2) quasinilpotent if r(T) = = 0.
Clearly, every nilpotent operator is quasinilpotent. Also notice that an
operator T E (X) is quasinilpotent if and only if a(T) = {0}.
Recall that if T: X Y is a bounded operator between Banach spaces,
then its adjoint operator T*: is defined via the duality formula
(x,T*y*) = (Tx,y*)
for all x E X and all E It is well known that the mapping T
from (X, Y) to (Y*, X*), is a linear isometry.
Theorem 6.14. The spectrum of an operator T E (X) coincides with the
spectrum of its adjoint, i.e.,
cr(T) = cr(T*).
In particular, we have r(T) = r(T*).
1
we use the fact that if A, B, C E (X) satisfy CA = BC = I, then A = B = C'.
To see this, note that A = IA = (BC)A = B(CA) = BI = B.
6.1. The Spectrum of an Operator 245
Exercises
=
dATh
= {A e A T is not one-to-one }
= {AEu(T):
Recall that if an operator is one-to-one, then (in view of the Open Map-
ping Theorem) it fails to be invertible if and only if it is not surjective.
Accordingly, when a noninvertible operator A T is one-to-one, we distin-
guish two cases:
(1) The range of the operator AT is not dense, i.e., (A T)(X)
(2) The operator A T has dense range, i.e., (A T)(X) = X.
In case (1) we say that A belongs to the residual spectrum of T.
That is, the residual spectrum is defined by
= {A e \ (A - T) (X) x}.
6.2. Special Points of the Spectrum 249
{A E a(T) \ (A T)(X) = X
Clearly, (T), and
o"(T) = U U
= {A E C: c X with I V n and -4 0
}
= {A E C: X satisfying 74 0 and 0 }
If AT does not have dense range, then by the Separation Theorem there
exists some e satisfying ((A T)x,x*) = 0 for all x e X.
But then (*) implies (A T*)x* = 0 or T*x* = Ax*, i.e., A is an eigenvalue
of T*. Conversely, if T*x* = Ax* holds for some e X*, then (*)
implies ((A T)x, x*) = 0 for all x e X, and so the range of A T cannot
be dense in X. D
Regarding eigenvalues we have the following interesting result.
Theorem 6.20. If an operator T: X X on a complex Banach space is
surjective but not invertible, then there exists some > 0 such that every A
in the open disk D(0, E) is an eigenvalue ofT.
Proof. According to Corollary 2.12 there exists some E > 0 such that every
operator S e (X) that satisfies < E is surjective and non-invertible.
In particular, for each A e D(0, E) the operator T A is surjective but not
invertible. This implies that each A e D(0, ) is an eigenvalue of T.
6.2. Special Points of the Spectrum 251
Let us use the shift operator as an example to illustrate the various parts
of the spectrum. The (forward) shift operator T: 4 (1 <p < oc)
is defined by
T(zi, Z2,...) = (0, zi, Z2,...).
For 1 p <oc we have = eq, where + = 1, and the duality eq)
is defined by
(z,e) =
T*(zi,z2,...)=(z2,z3,...),
o.(T)=o.(T*)={AEC:
and
crc(T*) = a(T*) \ [crp(T*) U Ur(T*)] E C: = i}.
=
4. The approximate point spectra of the forward and backward shift
tors.
Here we claim
Da(T) = i} c cra(T).
E C:
To establish that equality holds, now observe that for a! < 1 and x E 4
with lix lip = 1, we have
TxIIp = I
I
x
sj jT
x
Figure 1
is an eigenvalue of TS.
(b) If X is finite dimensional, then show that cr(ST) = a(TS); compare
thiswith Exercise 19 of Section 6.1 and Exercise 7 of Section
[HINT: (ST)x = Ax implies (TS)Tx = ATx.1
are equivalent, the following result should be immediate from Lemma 1.7.
Lemma 6.22. If T: E p E is a bounded operator on a real Banach lattice,
then
1 1
r(T) max Al = lim ]n lim
C
(2) If A1 and A2 are real numbers and A1 > A2 > r(T), then
o R(A2,T).
R(A1,T)
Proof. (1) Using Lemma and Exercise 1 at the end of the section,
see that
00 00
=
Assume that A is real and satisfies A > r(T). Then each operator
(2)
So, R(A, T) a positive operator. The other inequality now follows from
is
the resolvent identity by observing that
R(A2,T) R(A1,T) = (Ai A2)R(A1,T)R(A2,T) 0.
6.3. The Resolvent of a Positive Operator - 255
0 TR(A, T) >A(n+l)Tn+l =
Proof. Consider the resolvent operator R(A, T). Since A> r(T), we know
that R(A, T) = We claim that the vector u = R(A, T)v has the
desired properties. Indeed,
00 0000
Applying Lemma 6.23(3), we get
Tu = TR(A,T)v AR(A,T)v = Au.
Finally, from R(A,T) we obtain u = R(A,T)v This implies
Exercises
1. Assume that exists in E, where = + e for each n.
Show that exists in and 1 where the
series are assumed to converge in norm. [HINT: If k and 9 are fixed, then
f f(A)R(A, T) dA
= f T)
f(T)=
2 means that there exists some neighborhood U of cr(T) such that C1 U and f
analytic on U.
6.4. Functional Calculus 259
+fig(T)
= f
= + T) = + fig)(T).
and
E C2 dA = 0 for each E C1. Therefore, using Fubini's
theorem and the Resolvent Identity (see Definition 6.5), we see that
f(T)g(T)
=
f
C2 f
C1
: EL2
T)
T) = (fg)(T).
1C2
The above establish that the mapping f 'k 1(T), from 1'(T) to (X), is
algebraic homomorphism. We shall finish the proof by verifying the
remaining properties.
(1) From ST = TS and Lemma 6.4 we see that SR(1\, T) = R(1\, T)S for
each e p(T). Now the commutativity of S and f(T) follows immediately
from the definition of the Riemann integral of a Banach-valued function.
(2) Theorem 6.14 asserts that o-(T) = o(T*), and so =
Also, we know that T*) = T)]* for all A E p(T); see Exercise 4 of
Section 6.1. Now the desired conclusion follows from the definition of the
Riemann integral by observing that:
f(T*)
=
=
260 6. Spectral Properties
h(T) =
f AkR(A T) dA =
f dA
This conclusion combined with the first part shows that for every polynomial
p(A) = we have p(T) =
We now consider the general case. Since the series f(A) = E=O
a neighborhood of o-(T), there exist 0 < ri < such that
cr(T) B(O, ri) and the series converges for each A E r2).
In particular, the series f(A) = converges uniformly on B(0, r1).
That is, if pk(A) = then the sequence of polynomials {Pk}
verges uniformly to f on B (0, ri). This implies (see Exercise 5 at the end
of the section) that
I.
We are now ready to state and prove the Spectral Mapping Theorem.
Theorem 6.31 (The Spectral Mapping Theorem). If T E (X), them for
every fumction f E we have cr(f(T)) = f(a(T)). That is,
a(f(T)) = {f(A): AE
f
Functional Calculus 261
= if E = E
and cr(TIza)=cr(T)\cr,
where and denote the restrictions oi T to Y0. and Z0., respectively.
= (T) + (T) = I.
This shows that the operator T is invertible on X or
a(T), contrary
to the assumption that E a a(T). Thus, ..\ a(T0.), and so a(T0.) = a.
Finally, we shall prove the uniqueness of the reducing pair. To this end,
assume that (Y Z) is another reducing pair for T satisfying
a and a Jordan
contour in surrounding and C2 a Jordan contour
in V0.' surrounding a(T). Now note that if y E Y, then we have
f =
This implies y E Y0., and so Y Y0.. Similarly, Z c Z0.. If v Y0., then
writev=y+zEYEDZ=X, andnotethatz=vy EY0.nZ0.={O}.
That is, z = 0 or v y E Y. Thus, Y0. Y, and hence Y = Similarly,
Z= and the uniqueness of the reducing pair has been established.
264 -- - - -- 6. Spectral Properties
is the La'urent series expansion of the resolvent function R(., T) aro'und A0,
then the coefficients of the expansion satisfy the identities:
A_1 = I+(TAo)Ao,
= (T for each n 0, and
= for each n.
In partic'ular, for each n 1 we have
(T = A0 and = (T
Proof. Multiplying R(A, T) = A T
and using that (A T)R(A, T) = I, we obtain
I= (AT)CAAo)mAn
fl 00
00 00
= + T)(A AO)71A7.,
= + (Ao
The uniqueness of the Laurent series expansion yields I = A_1 + (A0 T)Ao
and + (A0 T)A72 = 0 for each n 0. These are the first two desired
identities.
If we multiply R(A, T) = A
then as above we can obtain that = for each n. The last two
identities are straightforward.
If A0 is an isolated point of the spectrum, then the set a = {A0} is clearly
a spectral set. For simplicity we shall write P\0 (T) rather than (T).
In this case, we shall also write and instead of the more laborious
symbols and
266 6. Spectral Properties
around A0 the residue operator A_1 is a projection that coincides with the
spectral projection associated with {Ao}. That is, A_1 = PA0(T).
=
f dA for n =0, 1, 2
In particular, we have
A1 =
For the rest of the proof the symbol C5 will denote the positively oriented
circle centered at A0 and having radius 8 > 0. Now consider the open set
V = D(Ao, r) U {z E C: 2r < 1z Aol <p}, where p > 0 is large enough so
that u(T) V. Clearly, 3V = Cr U Cp U (C2r), where denotes the
circle oriented negatively.
Let f be the function defined by
f(A)cf0 if
if
Clearly, f belongs to 2(T) and is identically 1 on the open disk D(A0, r).
So, by the definition of the spectral projection P,\0 (T) associated with the
spectral set {Ao}, we see that f(T) = PA0(T). It remains to notice that
A_1 =
f R(A, T) dA = f(A)R(A, T) dA = f(T) = PA0(T).
Proof. (1) (2) We have A_k 0 and = 0 for all n > k. But
then it follows from Lemma 6.37 that (T = 0 and
(T A0)kA_1 = A_(k+1) = 0.
(2) (3) From Lemma 6.38, we know that A_1 = PA0(T). So, we
have (T 0 and (T = 0.
We shall verify first that N((Ao = N((A0 We already
know that N ((Ao C N ((Ao To prove the reverse inclusion,
let x E N((Ao and assume by way of contradiction that x does
not belong to that is, the vector y = (Ao 0. It
follows that (Ao T)y = (Ao = 0 or Ty = Aoy. This implies
(see Exercise 3 at the end of this section) that f(T)y = f(,\o)y for each
f In particular, letting f = f0., where f0. is the function defined
before Theorem 6.34, we obtain that = y. Consequently,
0 = (T Ao)kPAO(T)x = PA0(T)(T Ao)kx = = y 0,
which is a contradiction. Hence, N((Ao T)k) = N((Ao This
shows that the operator A0 T has finite ascent and that p = a(Ao T) <k.
In particular, we have N((A0 = N((Ao for all n p.
Next, notice that (T 0 guarantees the existence of
some vector x in the range of the projection PA0(T) such that
(T = (T 0.
From (T Ao)kx = (T = 0, it follows that
- N((Ao -
This shows that p k is also true. Thus, p= k.
Next, we consider the unique decomposition X =YA0 ZA0 described
in Theorem 6.34, where ZA0) is a reducing pair for T. ) = {A0}
and cr(TIZA ) = cr(T) \ {Ao}. In particular, the operator A0 T is invertible
on ZA0. This implies that for each n E N the operator (A0 T)Th is also
invertible on ZA0.
268 6. Spectral Properties
Exercises
1. Show directly (LeO, by avoiding the Spectral Mapping Theorem) that
if p(z) = a polynomial, then for every bounded operator
T: X X on a Banach space we have
=
Now notice that ji e a(p(T)) if and only if E (7(T) for some
2. Assume that f is an analytic function defined on a neighborhood V of CX),
that is, the function f is defined and is analytic on an open set of the form
V = {A e C: > R} for some R> Suppose also that f vanishes
at in V surrounding the spectrum of an
operator T E (X), then show that Jc f(A)R(A, T) dA = 0.
3. Let T: X X be a bounded operator on a Banach space. Fix any
function f e and let C be a Jordan contour surrounding cr(T) in
the domain of analyticity of f. Show that
f(T)x =
f f(A)R(A, T)x dA
15. (R. Melton) For an entire function f: C f C having the series expansion
f (A) the following statements are equivalent.
(a) Each coefficient of the expansion is non-negative.
(b) For each positive operator T on any Banach lattice E the operator
f(T) = is positive.
(c) There exists an infinite dimensional Banach lattice E such that f(K)
is positive for each positive compact operator K on E.
Chapter 7
271
272 7. Some Special Spectra
= E Va_i.
=
7.1. The Spectrum of a Compact Operator 273
This shows that {Tyn} cannot have any convergent subsequence, contrary
to the compactness of T. This completes the proof.
Our next goal is to show that each non-zero boundary point of the
spectrum is an eigenvalue, i.e., each e 9a(T) \ {0} belongs to
So, let,\ e 9a(T) \ {0}. By Theorem 6.18(2), we know that belongs to
the approximate point spectrum of T. So, there exists a sequence of
unit vectors such that T guarantees
that some subsequence of {TXn} (which we shall denote by {TXn} again)
converges. From = we see that the sequence {Xn}
is also norm convergent, say x. Since Ifr'nM = 1 for each n, we infer
Proof. The conclusions follow immediately from Theorem 7.3 and the fact
that = A E cr(T)}. (Use here either the Spectral Mapping
Theorem 6.31 or Exercise 11 in Section 6.1.)
Corollary 7.5. If T E L(X) is a compact operator, then each non-zero
A0 E cr(T) is a pole of the resolvent of T.
lo if 0<t<s<1
K(s,t)=<
if
and let V: 1] 1] (1 p oc) be the integral operator with
kernel K. That is, for each x E 1] we have
r1 rt
Vx(t)= / K(s,t)x(s)ds= / x(s)ds.
Jo Jo
This integral operator is known as the Volterra operator. According to
Theorem 5.29 its adjoint operator V*: Lq{O, 1] Lq[O, 1] is given by
fl I'1
V*y(s)=J y(t)dt,
0 S
where 1 q oc satisfies + = 1.
276 7. Some Special Spectra
Vx(t1) - Vx(t2)[ = f
Jt1
It1
1
for all t1, t2 E [0, 1]. This implies that { Vx: x I } is a bounded and
equicontinuous subset of C[0, 1], and from this we infer that V is a compact
operator.
The spectrum of V consists of the zero element alone, i.e., a(V) = {0}.
The fact that 0 E a(V) follows immediately from Theorem 7.1. On the
other hand, if 0, then for every y E 1], it is easy to check that the
equation
(\V)x=y
has a unique solution x (see Exercise 3 at the end of the section) given by
x(s)
This shows that V)' exists for each 0, and therefore a(V) = {O},
V is a quasinilpotent operator. Finally, it should be noticed that 0 is
not an eigenvalue of V.
= r(T)]xn =
we obtain that
1 2 /
+
By the compactness of T, the sequence has a convergent subsequence,
which we shall denote by again. From r(T) > 0 and (*), it follows
that x for some positive unit vector x. Using (*) once again, we obtain
Tx = r(T)x, as desired.
Lemma 4.55, there exists some open disk centered at and such that
for each ,a E the operator ,a T is Fredholm and its nullity n(,a T)
and defect d(,a T) are constant on the set \ Moreover, we claim
that T) = d(,a T) = 0 on with the possible exception of the
point
To establish this claim, fix some 0 and then choose any point
with > and such that the closed line segment L = joining
and does not include zero. Since the operator ,a T is invertible whenever
11TH, it follows that n(,a T) = d(,a T) = 0 for all ,a close to and
consequently for all ,a E \ by our choice of the disk
Exercises
1. Show that the only compact operators with closed range are the finite-
rank operators.
2. Consider the interval [0, 1] equipped with its Lebesgue measure and for
each 1 p oo define the positive integral operator T: 1] + 1]
by
f(s) =
f + kg(s).
Sf(t) =
f +
4. Show directly (i.e., without using the previous exercise) that the Volterra
operator has no eigenvaluesand conclude from this that the Volterra
operator is quasinilpotent.
5. Let T: E + E be a positive contraction on a Banach 'attice. Show that
1 belongs to the approximate point spectrum of T if and only if r(T) = 1.
6. Show that the hypothesis r(T) > 0 cannot be dropped from the Krein
Rutman Theorem 7.10.
7. Show that the compactness hypothesis of the operator in the Krein
Rutman Theorem 7.10 cannot be dropped. [HINT: Use the shift operator
in Example 6.21.]
8. Let T: E * E be a positive operator on a Banach lattice and let p be a
polynomial with non-negative coefficients. Show that r(p(T)) = p(r(T)).
[HINT: Use Exercise 1 of Section
9. Prove the following generalization of the KreinRutman Theorem 7.10.
If a power compact positive operator T: E * E satisfies r(T) > 0, then
there exists some vector x > 0 such that Tx = r(T)x. [HINT: Assume
that r = r(T) > 0 and that Tk is compact for some k> 1. According to
Theorem 7.10 there exists some x > .0 such that Tkx = rCx. Now notice
that if S = then y = Sx> 0 and
Tyry=TSx--rSx=(Tr)Sx= (Tic _rc)x=O,
proving that Ty = ry.]
10. Prove Lemma 4.18 using the KreinRutman theorem. That is, show by
using the KreinRutman theorem that if ci is a compact Hausdorif space
without isolated points, then the only weakly compact multiplication op-
erator on C(ci) is the zero operator. [HINT: Assume that there exists
a non-zero weakly compact multiplication operator on C(ci). Then
= is a non-zero compact positive multiplication operator;
see [30, p. 337]. So, replacing q5 by q52, we can suppose that q5> 0 and
that is compact. Clearly,
1 1
r = = lim
1
= = >
(Xi,x2,...) EHXn:
We also define
limIlxnII=O}.
fl*oO
n=1
It is also customary to denote these Banach spaces by 4(Xi
and co(Xi X2 E13 ) In the case X = = we shall write
and
=
the quotient Banach space of over the closed subspace co(X).
Our next objective is to show that the Banach space X has the properties
stated at the beginning of the section. -
= tI[(Txi,Tx2, . = limsup
< limsup
1 1
T is that Now if
x E X satisfies < 1, then = lxii 1, and so 11Th = ITxM.
This implies
ITII sup 1Txl = tTII
lix
We are now ready to establish the desired the relationships between the
spectra of the operators T and T. We start with the complex version of this
result.
Theorem 7.16. If T: X X is a bounded operator on a complex Banach
space, then:
(1) p(T) and =
(2) = = cTp(
(3) R(A,T) = R(A,T) holds for each A E p(T).
Proof. (1) If A E C, then it follows from Theorems 7.14 and 7.15 that the
operator A T is invertible in L1(X) if and only if A T is invertible in (X).
This shows that p(T) = p(T), and so o(T) = C \ p(T) = C \ p(T) = o(T).
(2) To prove that 0a(T) = notice that:
AE 0a(T), and let > 0. This implies that there exists some
vector [z] = [(zi, z2, .)] E X satisfying
1
[z] = lirn sup =I and I
= lim sup
Th-400
Then there exists some k such that < holds for all n k. Since
= 1, there are infinitely many m k such that >
This shows that A E 0a(T), and so 0a(T) Since 0a(T) 0a(T)
obviously true, we see that 0a(T) = c7a(T).
7.2. Turning EigenvaJues into EigenvaJues 285
- =
and the proof is finished.
(1) a(T)=cr(Tu).
(2) cra(T) = =
Exercises -
(xx*)
isasurjective
(topological) isomorphismwhich is a lattice isomorphism if X is
also a Banach lattice.
[HINT: For the continuity of T note that
fT({x + = 11[xI + = sup cos9 + [y] sin 911
9ER
= sup 1 [(xi cos 9 + sin 9, cos 9 + Y2 sin 9,. .
= IL[x+zy]IJ.]
00
IT I
Figure 1
subgroup of I' is a cyclic set. Also, according to this definition, every subset
of real numbers considered as a set of complex numbers is a
cyclic set.
H. B. Schaefer [290, Proposition 4.2, p. 324] has shown that the periph-
eral spectrum of a Markov operator enjoys the following "cyclicity" property.
Theorem 7.20 (Schaefer). Let E be an AM-space with unit e, and let
T: E E be a Markov operator. Then the set of eigenvalues
E={AET: TcZ=AZforsomeZEEc with IZI=e}
is a subgroup of the unit circle C (and hence E is a cyclic set).
Proof. Since Te = e, we have 1 E E. By Theorem 3.20, we can assume that
E = C(11) and = for some compact Bausdorif space where
the unit e corresponds to the constant function 1 on ft Also, we know
that E* = ca and = ca zca where ca is the
of all regular (finite) Borel measures on Since T is a Markov operator,
T* carries regular Borel probability measures to regular Borel probability
measures. For simplicity, we shall denote by T.
Now ffx w E and let = T*&,. Assume that there exist Z E
andAETsuchthat ZJ=landTz=Az. Thenwehave
and so f
AZ(w) = (TZ)(w) = (TZ, = (Z, = (Z, /2w)
=
= 1. Since P\'[Z(w)]'Z(t)I =
f Z(t)
1 for each
t E it follows (how?) that ?C'Z(t) [Z(w)j' = 1 for each t lying in the
support of This implies Z(t) = AZ(w) for all t in the support of
Similarly, if another E I' satisfies TZ1 = A1Z1 for some E
with JZi 1 1, then Zi(t) = A1Z1(w) for all t in the support of and so
= for all tin the support of Hence,
AA1Z(W)Zi(W)
= (TZZ1,
= f Z(t)Zi(t)
= (TZZ1)(w).
= = T*8W)
Now if we fix some real number A0 > r(T), then according to Lemma 6.24
there exists some 0 < q5 E E* such that the principal ideal generated by q5
in E* is E and A0q5. By Theorem 3.39,
there exists a unique lattice homomorphism Tgt,: + whose adjoint
makes the diagram in Figure 2 commutative. (The construction of the
space is described in Section 3.4.)
290 7. Some Special Spectra
Figure 2
From the construction of the space and the fact that T is a lattice
homomorphism, for each x E E we have
AJ =
=ci(TIxI)
= =
=
This implies for each z E i.e., is a bounded
below operator.
Now observe that if 0 then T is a surjective lattice isomorphism,
and since is a lattice isometry, T*: is also a lattice isomorphism.
Since A E it follows from Corollary 7.22 that
E o(T)
for each k E Z.So, in this case, A is a cyclic point of the spectrum cr(T).
Next, assume that 0 E Since T17!, is bounded below, it follows (see
Exercises 5, 6, and 8 of Section 6.2) that
D(O, Aj) = cYp(T*JE;) =
That is, the open disk centered at zero with radius Al lies in o(T), and so
A is automatically a cyclic point of o(T), and the proof is finished.
Corollary 7.24. The spectrum of an interval preserving positive operator
is cyclic.
Exercises
G= {z C: zP = 1} = k
r0(T) lim
(See Theorem 6.16 and the discussion preceding it.) It should be clear that
for any bounded operator T we have the inclusion o(T) a0(T).
In what follows, we will restrict our attention to order bounded oper-
ators. Let us mention at once that in general the order spectrum a0(T)
can properly include the usual spectrum a(T) (see Example 7.36 below).
Nevertheless, there exists a method that allows us to reduce the notion of
the order spectrum to the usual spectrum. More precisely, each Dedekind
complete Banach lattice E can be embedded into a larger Dedekind com-
plete Banach lattice E in such a way that each order bounded operator T
on E will have an order bounded extension D to this enlargement satisfying
= a(T). The purpose of this section is to describe this method.
We begin by assuming that E is a Riesz space and that A is an arbitrary
fixed index set. Only this is necessary to define the extension E.
It is well known, and easy to check, that the Cartesian product EA under
the pointwise operations is also a Riesz space. That is, EA is a Riesz space
under the operations defined by
[f + g}(a') + g(a), = and [f V
g all E E A. The positive cone of EA consists
of all positive functions, that is, functions f E EA with E for each
Lemma 7.26. A Riesz space E is Dedekind complete if and only if the Riesz
space EA is Dedekind complete.
= T(f(a)).
where f Notice that T is linear if and only if T is linear.
E A.
The reader should verify also that if S, T: E E are two operators and A
is an arbitrary scalar, then
(S + = + t, = At, and =
Regarding invertibility of operators we have the following.
Lemma 7.29. Let T: E p E be an operator on a Riesz space. Then:
(1) T is one-to-one if and only if T: E E is one-to-one.
(2) T is snrjective if and only if 1': E'1 EA is snrjective.
294 7. Some Special Spectra
Proof. (1) Assume that the operator T is one-to-one and let Tf = 0. This
implies = T(f(a)) = 0 for each E A. Since T is one-to-one, it
follows that = 0 for each a E A, f = 0. Rence, is one-to--one.
Now suppose that T is one-to--one on E, and let Tx 0. So, = 0,
and thus = 0. This implies x = 0, and hence T is one-to-one.
(2) Assume that T: EA pjA
is surjective, and let y E E. Pick some
f E EA such that Tf = and then note that for each index E A we have
= T is surjective.
For the converse, assume that T is surjective, and let f E EA. For each
a E A pick some E E such that = Then the function
gE EA satisfies Tg = f, so that is surjective.
= V = V T(x V y).
This shows that T is a lattice homomorphism.
(3) Assume that T is interval preserving, and let f 0 and g E
satisfy 0 g Tf. That is, 0 = holds for each
a' E A. ,Since T is interval preserving, for each E A there exists some
7A. The Order Spectrum of an Order Bounded Operator 295
Proof. Assume first that T is order bounded, and let f E E. Pick some
xE satisfying x for each a E A. Since T is order bounded,
there exists some y E such that T[x,x] [y,y]. Now note that
Tf(a)1 = y for each a E A. Therefore, Tf E E, and so T
leaves E invariant. Moreover, notice that for each g E E satisfying f
we have Tg This shows that T: E E is indeed an order
bounded operator.
For the converse, assume that T leaves E invariant. Fix some x E E+.
Since the cardinality of the index set A is at least as large as the cardinality
of E, there exists a function f: A E whose range is the entire order
interval [x, x]. In view of our hypothesis, Tf E E and so there exists an
element y E such that Tf I Therefore T[x,x] [y,y}. This
proves that T is order bounded, and the proof is finished.
part (2) of the same lemma shows that T is also surjective. Consequently,
T' existsand, of course, is a linear operator.
From TT1 = T'T = I. it follows that = = I on
Ek Since 1' leaves E invariant, it follows that S = (T)' = (T1)"; see
Exercise 1 at the end of this section. Finally, observing that (T)' leaves
E invariant, we from Theorem 7.31 that T' is also an order bounded
operator.
Recall that every linear operator on E zE is of the form S + iT,
where S and T are linear operators on E. Let us say that an operator S +zT
is order bounded if both S and T are order bounded. Also, if S + iT is an
operator on let us write (S + + zt; clearly, this operator acts
on (EA)c. Notice that if S and T are order bounded, then S + iT leaves the
vector subspace = E zE of (EA)c invariant. The complex version of
Theorem 7.32 is now immediate.
Theorem 7.33. Let 5, T: E E be two order bounded operators on a
Riesz space. If (S + zT)": i is invertible, then S + iT: i is
also invertible, and its inverse is order bounded and satisfies
[(S + [(S + = +
Theorem 7.33 implies at once the result announced at the beginning of
the section about reducing the order spectrum to the standard spectrum.
Theorem 734. If T is an order bounded operator on a Dedekind complete
Banach lattice, then cr0(T) = = cr(T).
Corollary 7.35. Lattice homomorphisms and interval preserving positive
operators have cyclic order spectra.
Proof. Let T: E i E be a positive operator which is either interval pre
serving or a lattice homomorphism. According to Theorem 7.34 we have
cr0(T) = cr0(T) = cr(t). Also, from Lemma 7.30, we know that T: E E is
either interval preserving or a lattice homomorphism, respectively. To com-
plete the proof, it remains to notice that by Theorem 7.23 and Corollary 7.24
these types of positive operators have cyclic spectra.
We conclude the section with the promised example, due to T. And
(see [291]), of an order bounded operator for which the order spectrum is
larger than the usual spectrum.
Example 7.36 (And). For each k consider the finite dimensional real
Hilbert space Ek = of Ic factors of with the stan-
dard coordinatewise inner product. With the usual coordinatewise order
becomes a Dedekind complete Banach lattice.
7.4. The Order Spectrum of an Order Bounded Operator 297
(x,y) =
for each m E N.
I
cost
Therefore, the modulus of (considered as an operator on R2) is given
by the 2 x 2 symmetric matrix
cos f
sin
k
It is easy to see that the eigenvalues of are f
cos + sin So, the
Eucidean norm of is = cos + sin j.
f
Next, for each k E N we define the operator Tk: Ek Ek via the formula
Tk = Mk . Clearly, Tk is a unitary operator on Ek which is
also order bounded. This implies fITkII = 1. Moreover, the modulus of
is given by ED jA'Ij E
Fix any 0 < < I and then pick some no such that cos > and
sin > for each n no. This and Exercise 4 in Section 7.2 imply that
for alln no we have
= sup
kN
=
>
=
298 7. Some Special Spectra
Consequently, we have
r0(T) = n*oc
lim T)
lim [(i + = e> 1= r(T).
So, the spectrum of T is a proper subset of the order spectrum of T.
Exercises
1. Let T: V V be an invertible operator on a vector space, and let
V V be its inverse. Assume that W is a T-invariant subspace of
V such that T: W TV is also invertible. Then show that T' leaves TV
invariant and that the inverse of Tjw (the restriction of T to TV) is the
restriction of T' to W, i.e., (Tlw)1 T11w.
2. Describe the Banach lattice B when B = R.
3. Prove Lemma 7.26.
4. Prove Lemma 7.28.
5. Prove Theorem 7.33.
6. Prove Theorem 7.34.
7. Establish the following elementary properties that were used in Exam-
ple 7.36. Consider the matrix
o [S]+[T]=[S+T],
o A{T] = [AT], and
o [S][T] = [STJ
is a unital algebra with the element [I] as its unit. Moreover, the quotient
vector space L(X)/IC(X) under the quotient norm
= S -T e
is a Banach space. Since the quotient norm also satisfies the properties
and 1['IM 1,
L(X)/IC(X) is a unital Banach algebra. This Banach algebra is called the
Calkin algebra' of X and is denoted i.e., = L(X)/IC(X).
The quotient map (also called the canonical projection) of (X) onto
will be denoted by ir. That is, it: (X) is defined by
The next result identifies a special class of isolated points in the spectrum
of a bounded operator.
Lemma 743. Let T E (X) and )'o E \ If there is a path
lying outside of ess (T) and joining with a point in the resolvent set p (T),
then is an isolated point of the spectrum o(T).
Proof. Fix E \ and let C be a path joining with a point
E p(T) such that CflO'ess(T) = 0. So, for each z E C the operator zT is
a Fredholrn operator. Now, according to Lemma 4.55, for each z E C there
exists an open disk D(z) centered at z such that the nullity n(( T) and
the defect T) are constant on D(z) \ {z}.
in the open disk D(zi) with the exception of ( = This implies that T
is invertible for all ( \ So, is an isolated point of o(T).
Theorem 7.44. Let be an isolated point in the spectrum of an operator
TEL(X). If then:
(1) The spectral projection P,\0(T) is of finite-rank.
(2) The operator )'o T has finite ascent and descent.
(3) The point is an eigenvalue of T and is a pole of the resolvent of
T whose order equals the ascent (or the descent)
X=X1EBX2 -T1)eYeX2
Since T is Fredholm, Y is finite dimensional, and so from
(2) and (3) Since X1 is finite dimensional, it follows that \oix1 T1 has
finite ascent and descent. This, in conjunction with (*) and (**), guaran-
tees that T likewise has finite ascent and descent. Now a glance at
Theorem shows that is a pole of the resolvent of T.
Since cr(T1) = {\o} and dim X1 <oo, there exists some non-zero vector
x E Xi such that Tx = T1x = Thus, is an eigenvalue of T. The last
conclusion follows immediately from Theorem o
302 7. Some Special Spectra
Proof. The implication (b) (a) has been established in Theorem 7.44.
Now assume that (a) is true.
Let Xi = (T)) and X2 = N(PA0 (T)). to our hypothesis
According
X1 is finite dimensional, X = X1 e X2, and the closed subspaces X1 and
X2 of X are both invariant under T. If T to then
cr(T1) = and cr(T2) = cr(T) \ Since cr(T2), it follows that
T2 E (X2) is an isomorphism.
Fredholm operator.
We now introduce the essential spectral radius of a bounded operator.
Definition 7.46. The essential spectral radius of an operator T E (X)
is the spectral radius of the element ir(T) in the Calkin algebra i.e,
ress(T) r(ir(T))
= E cress(T)}
= lim
flP 00
=
inf jJ TTh j*.
The essentially quasinilpotent operators are defined as follows.
Definition 7.47. An operator T E (X) is said to be essentially
nilpotent (or a Riesz operator) if aess(T) = {0}, or equivalently, if its
essential spectral radius is zero.
The next result presents a useful characterization of the essentially quasi-
nilpotent operators.
Theorem 7.48. An operator T E (X) is essentially quasinilpotent if and
only if every E a(T) \ {0} is:
(1) an isolated point of a(T), and
For the converse, assume that every point E cr(T) \ {O} satisfies (1)
and (2). Now fix E u(T) \ {0}. By (1), we know that is an isolated
(b) \ is an eigenvalue of T,
(c) the operator T has finite ascent and descent, and
(d) the order of the pole equals the ascent (or the descent) of the
operator T.
and the definition of the Riemann integral, we infer that each is a com-
pact operator. In particular, we have = for
each n 0. From = 0, we get
for each P'I > r(T), it follows that the series = converges
for each .AJ > r(T) to a (necessarily) compact operator. In particular, we
have R(.A, T) B(\) + K(,\) for each > r(T).
The function K: p(T) > (X) defined by =
is clearly analytic. Since = for each > r(T) and p(T) is
connected, it follows that K is the unique analytic extension of K to p(T).
To finish the proof, it remains to be shown that is a compact operator
e p(T). That is, is a compact operator for each e p(T), and the
proof is finished.
Our next goal is to establish that the Calldn algebra can be represented
as a subalgebra of bounded operators on an appropriate Banach space. To
do this, we need some preliminary discussion.
Let U be an ultrafilter on N and let X11, be the ultrapower of X deter-
mined by the ultrafilter U; see Section 1.4. If T E L(X), then there exists a
natural extension of T to an operator T11, on X11, defined by
TU[(Xi,X2,. [(Tx1, Tx2,.. = (Tx1, Tx2,...) +
Let us verify that the above formula is well defined and that it defines a
bounded operator on X. Fix E [xv]. This means that there exists some
u E X such that u 0. Consequently, Tn 0, and
so (Ty1, Ty2, E [(Ty1, Ty2, . .)u]. This shows that T is well defined,
.
=
= .)u]M
S ITuII
and from this we easily get
(
That is, T: X X is indeed a bounded operator.
The basic algebraic properties of the operators T are included in the
next result. These properties (as well as many others) have been obtained
by J. J. Buoni, R. Harte, and A. W. Wickstead [79], J. J. Chadwick and
A. W. Wickstead [87], and M. P. W. Wolff [340].
Lemma 7.54. For an operator T E (X) we have the following.
(1) T = 0 if and only if T is a compact operator.
(2) T is one-to-one if and only if T is a semi-Fredhoim operator with
finite nullity, i.e., n(T) < 00. -
Proof. (1) Fix an operator T E C(X). Assume first that T = 0 and let
{ be a sequence in If we let x = (xi, x2,...) E then from
= [(Tx1, Tx2,. . = 0, it follows that there exists some z E X such
that z. In particular, it follows (see part (3) of Lemma 1.59) that
there exists a subsequence of such that p z in X. This
(*)
308 7. Some Special Spectra
aM
<1 (**)
< +
= II
< aM +
= 0. Next, notice that (in view of Corollary 2.15) there exists a sequence
x = (xi, x2,...) E (X) such that = for each m. Finally, notice
that the identities = + are equivalent to = + This
implies
= = + = Tx + = Tx,
and tlie proof of this part is finished.
(4) This follows immediately from properties (2) and (3)
that for each n the set A is not a subset of So, for each n
there exists some x E We
shall finish the proof by showing that =
To see this, start by observing that from Jjxn zkII r for each n> k
it follows that zkII = lirnu r, and so zkjl for
each k. Since Z is dense in X, we have 4 for each z E X.
But then we have
d(x,X)
That is, = as
=
=
= }
=
=
and the proof is finished.
= Tx2,. .)u] .
=
we obtain that ST = ST so that the mapping [T] 'p T is also an algebraic
homomorphism.
The fact that [TJ T is a linear isometry when is equipped with
the norm follows from Corollary 7.58. This guarantees that the operator
{T] T is also an algebraic isomorphism.
Next we will establish an interesting formula, due to R. D. Nuss-
baum [258], for calculating the essential spectral radius of a bounded oper-
ator.
Theorem--7.61--(Nussbaum). -For a bounded operator T: X X on a
Banach space we have the following.
(1) The spectrum of T coincides with the essential spectrum of T, i.e.,
a(T) aess(T).
(2) If X is separable, then the essential spectral radius of T satisfies
Proof. (1) Using part (4) of Lemma 7.54 and the definition of the essential
spectrum, we see that:
AE A T is not invertible
A T is not Fredholm
AEc.Tess(T).
312 7. Some Special Spectra
(2) This follows immediately from part (1) and Corollary 7.58.
T. T. West [333] has proven that the converse of the preceding conclusion
is true for Hubert spaces. We state this result below. Its proof is beyond
the scope of this monograph and is omitted.
Theorem 7.62 (West). A bounded operator T: H H on a Hubert space is
essentially quasinilpo tent if and only if there exist a quasinilpotent operator
Q E (H) and a compact operator K E i'C(H) such that T = Q + K.
We shall also say that a Banach space X has the West Decomposition
Property if every essentially quasinilpotent operator on X has a West
decomposition. In this terminology, Theorem 7.62 states that:
Every Hilbert space has the West Decomposition Property.
The following question arises here:
o Does every Banach space have the West Decomposition Property?
The answer to this question for general Banach spaces is unknown.
K. R. Davidson and D. A. Herrero [99] have shown that the Banach spaces
CO and 4 for 1 <p < 00 have the West decomposition property.
There is another special part of the spectrum of an operatorknown
as the Weyl spectrumthat is related to the essential spectrum and to the
West decomposition property. Its definition is as follows.
Definition 7.63. The Weyl spectrum of an operator T E (X) is
the intersection of the spectra of all compact perturbations of T, i. e.,
fl u(TK).
KEIC(X)
7.5. The Essential Spectrum ofa Bounded Operator 313
Clearly, c o-(T).
it is not known if there is a particular
compact operator K such that cr(T K) = That is:
Is the Weyl spectrum of a bounded operator equal to the spectrum
of a compact perturbation of the operator?
Exercises
1. For the identity operator I on a Banach space X establish the following:
(a) Xis infinite dimensional if and only if = 1.
(b) X is finite dimensional if and only if [I] = 0.
2. Let A be a unital Banach algebra, and let G(A) denote the non-empty
collection of all invertible elements of A. Show that G(A) is an open
subset of A and that the mapping x 'p x1, from G(A) to G(A), is
continuous. In particular, G(A) is a topological group. [HINT: Note that
if e is the unit element of the algebra A and if < 1, then the element
e a is invertible and (e a)' =
a that some x E A satisfies tx < 1a1IV Then
from Ie xa'tj IRa x)a'IL <1, it follows that
xa1 = e (e xa') is an invertible element. This implies that x is
invertible. For the continuity of x x1 see Exercise 11 of Section 2.1.]
3. Let 5: 4 where 1 p < Do, be the forward shift defined by
S(xi,x27x3,...) = (0,xI,X2,x3,...).
Show that the essential spectrum of S coincides with the unit circle, that
is,
Positive Matrices
315
316 8. Positive Matrices
(x,y) =
The unit sphere of C7' under the Euclidean norm will be denoted by S,
i.e., S = { x E = = 1 }. The positive part of S will be
denoted by i.e. {x E 5: x O}. Both S and are compact
subsets of Whenever necessary, 5+ will be considered as a compact
subset of RTh. Recall that a positive vector x > 0 in is said to be strictly
positive, in symbols x >> 0, whenever each component of x is positive.
We shall consider any given n x n matrix (with complex or real entries)
as an operator acting on via the usua' matrix multiplication formula.
Recall that throughout this book the vectors in C7' are usually denoted as
row vectors; for instance, we write x = (Xi, X2,.. . , However, whenever
we encounter the expression Ax, where A = is an n x n matrix, then
8J. The B4anach Lattices and 317
Ax= =
arnjxj
If n matrix
[ajj] is au withcomplexeuitries, A = [iT]
conjugate of A and At where = is the transpose of A.
Clearly, At = At. We have the fundamental duality ideuitities:
and
for-all x,y e
It is well known that each A E o(A) is an eigenvalue of the matrix A, that is,
cr(A) = If A1, A2,.. , are the roots of the characteristic polynomial
.
Exercises
1. If A is an n x n matrix with real or complex entries, then show that
t
(Ax,y) (x,A y)
= (Az,z) = (Az,z) =
is called diagonalizable if there exists an
A
invertible Ti X Ti matrix R (called a diagonalizing matrix for A) such
that the matrix R'AR is diagonal. Establish the following.
(a) A matrix A is diagonalizable if and only if A has Ti linearly indepen-
dent eigenvectors.
(b) Every Hermitian matrix is diagonalizable by a unitary matrix.
7. An Ti X Ti matrix A is said to be positive semidefinite if (Az, z) 0 for
all z E Ctm. Establish the following.
(a) If A is a positive semidefinite matrix, then A has non-negative eigen-
values.
(b) A Hermitian matrix A is positive semidefinite if and only if it has
non-negative eigenvalues.
[HINT: For (b) use the fact that a Hermitian matrix is diagonalizable by
a unitary matrix.]
320 8. Positive Matrices
IhAxll2
= (x,AtAX)
=
This implies hAll =
10. Show that a square matrix is quasinilpotent if and only if it is nilpotent.
(Compare this with Exercise 18 of Section 6.1.)
11. Let A = 6 If we let = + and consider the real
matrices B = and C = then we can write A B + iC and view
A as a vector in the complexification = of the
real Banach lattice
Show that the modulus of A = B + iC in is the
positive matrix Al = [IajjhI.
12. Prove the Cayley_Hamilton Theorem 8.3.
= =
A1
x= = = A, we have
Tx =
:1=1 j=1 i=1
ei) = E (E
i=1 j=1
In other words, for each fixed basis, the operator T can be identified
with an n x n matrix. In particular, we can identify (X) with (C).
This means that once a basis of X has been chosen, then every n x n
matrix definesas abovea unique operator on X.
If we consider another basis {fi, f2,.. fn} in X and let B be the repre-
. ,
(a)
satisfy properties (a), (b), and (c) for the operator T: X X, and the
proof is finished.
8.2. Operators on Finite Dimensional Spaces 323
independent.
Now assume that X is finite dimensional and that T is nilpotent with
index of nilpotence q. Hence, there exists at least one vector u E X such
that 0. This implies that q dim(X). D
2This is also expressed by saying that: Every operator on a finite dimensional vector space
is the direct sum of a nilpotent operator and an invertible operator.
324 8. Positive Matrices
(1) and
0 0 for each i.
(2) The collection of vectors
,. . TU2, u2,
.. Tub, uk
,
is a basis of X.
In particular, the matrix A representing T with respect to this basis has
the following form.' Every entry of A outside the diagonal above the main
diagonal of A is zero. On the diagonal above the main diagonal, we have
(starting from the top) a string of qi I ones followed by a zero if k 1.
Then there is a string of 1 ones followed by a zero if k 2. This pattern
follows until we exhaust all the qj.
plex vector space. Assume that )'i, .. are the distinct roots of the charac-
,
Proof. According to Theorem 8.5, for each i = 1. ... k there exists a pair of
,
(a) The entries on the main diagonal of the matrix B are the eigenval-
ues of A repeated according to their multiplicities. Starting from
the position (1, 1), we have the eigenvalue A1 repeated m14imes
followed by the eigenvalue A2 repeated m2-times, and so on.
(b) All entries of B which are not on the main diagonal or on 'the
diagonal above the main one are zero.
(c) On the diagonal above the main diagonal the entries are only zeros
and ones and have the following pattern: On the part above the
string of the Aj, the entries start with a zero followed
Exercises
1. Show that for an operator T: X X on a finite dimensional vector space
the following statements are equivalent.
(a) T is invertible.
(b) T is one-to-one.
(c) T is surjective.
[
HINT: By Exercise 2 above, we know that every matrix is similar to an
upper triangular matrix, and so we can suppose that A is itself an upper
triangular-matrix. In particular, the diagonal entries of A are the eigen-
values Aj; each appearing timesthe multiplicity of the eigenvalue
It follows that q(A) is also an upper triangular matrix having diagonal
entries q(A1),.. . , with each appearing rnj times.]
4. This exercise describes the minimal polynomial of an operator T: X X
on an n-dimensional complex Banach space.
(a) Show that the spectrum of T is finite and consists of
Moreover, if we let o(T) = {A1,... then show that k <
, n.
(b) Show that there exists a non-zero p(A) such that p(T) =0.
(c) Prove that some root of the polynomial p(A) in (b) must belong to
the spectrum of T; and so the spectrum of T is non-empty.
(d) Assume that 7r(A) is a polynomial of minimal degree and leading
coefficient one satisfying 7r(T) = 0. Show that
7r(A) =
where, as in (a), o-(T) = . . , and each is the ascent of
the operator T. (This establishes that the polynomial 7r(A) is
uniquely determined; it is referred to as the minimal polynomial
of the operator T.)
(e) Show that a q(A) satisfies q(T) = 0 if and on'y if q(A) is
divisible by the minimal po'ynomial 7r(A) of T. In particu'ar, prove
that two polynomials and satisfy ''(T) = if and only
if there exists a polynomial 9(A) such that = + 9(A)7r(A).
[HINT: Let {x1,.. , . be a basis of X. Since for each i the n + I vec-
tors .
, must be linearly dependent, there exists a non-zero
polynomia' such that = 0. If p(A) = p1(A)p2(A) . .
the result for the special case when S is a projection, Le, 82 = S. For
the general case show that there exists an invertible operator R: X X
such that RS is a projection, and then apply the preceding conclusion to
the pair of operators RS and TR'.
8. Let V be a finite dimensional vector space. If A is a non-empty subset of
V, then its annihilator A-'- is the vector subspace of V* defined by
v*(a)=O forall aEA}.
Similarly, the annihilator of any non-empty subset B of
V defined by B-'- = {v E IT: b*(v) = 0 for all b* E B}.
Assume that T: X Y is an operator between two finite dimensional
vector spaces. As usual, we define its adjoint operator T*: via
the duality identity
(Tx,y*) = (x,T*y*) = y*(Tx)
for all x E X and all E Establish the following:
(a) = Ker
(b) [Ker
(c) dim R(T*) = dim R(T).
9. If T: X Y is an operator between finite dimensional vector spaces,
then show that
dim Ker(T) + dim R(T) = dim X.
10. Use Theorem to present an alternate proof of the CayleyHamilton
Theorem
11. Show that every square matrix is similar to its transpose.
12. Let T: Y Y be a bounded operator on. a Banach space. A vector y E Y
is said to be cyclic (or more precisely T-cyclic) if the vector subspace
generated by {y, Ty, T2y, T3y, .} is dense in the Banach space Y.
.
with )tx>>Ax}
= {\0: with Ax>>Ax}.
Notice that if 0 A1. A and E LA, then A1 E LA and, similarly,
A1 and imply E UA. Notice also that if x>-O---
satisfy Ax Ax, then
= IIAxM hAil Mxli
and so )t hAil. That is, LA c [0, IAII]. Since 0 E LA, we see that LA is
a non-empty interval in R that is bounded from above. The set LA is also
closed. To see this, assume that a sequence c LA satisfies A. For
each ri pick some E with Since is a compact set, by
passing to a subsequence (if necessary), we can assume that x E S+.
Hence, Ax = A E LA and
so LA is closed. Consequently, LA = [0, AA], where
= sup LA = maXLA.
The set UA will be described in Corollary 8.23. The properties of
non-negative number AA will be discussed in detail below. The first property
follows immediately from the definition of AA.
Lemma 8.13. If two ri x ri positive matrices A and B satisfy A B, then
AAAB.
More properties of AA are included in the next result.
Lemma 8.14. For an n x ri positive matrix A we have the following:
(1) AA = r(A), the spectral radius of A.
(2)
A is strictly positive, then AA > 0.
(4) AA is the only possible eigenvalne of A having a strictly positive
eigenvector.
Proof. (1) By Theorem 8.11 there exists some x> 0 with Ax = r(A)x and
so r(A) <AA. On the other hand, if x E and A 0 satisfy Ax Ax,
Exercises
5. (Gershgorin [134]) This exercise presents a simple but very useful "ap-
proximation" of the spectrum of a square matrix. Let A = E (C)
and for each i let = and = {z E C: z <
Show that u(A) C2. [HINT: Let A be an eigenvalue of the ma-
trix A and pick x = 0 satisfying Ax = Ax. Choose
some k such that jxk = > 0. Now use the equation
(A akk)xk = to get IA
a E >
satisfies for each i, then
show that A is invertible. [HINT: Use the preceding exercise.
7. A non-negative n x n matrix A = is said to be a Markov matrix
whenever the sum of the elements in each row equals one, whenever
= 1 for each i. (This is equivalent, of course, to saying that A
defines a Markov operator on Wi.) Show in four different ways that every
Markov matrix A satisfies r(A) = 1 by using:
(a) Exercise 4 above,
(b) Lemma 8.14(4),
(c) Gershgorin's theorem (see Exercise 5 above), and
(d) the spectral radius formulaS
[HINT: If x = (1, 1,.. . , 1), then Ax = x, so that us an eigenvalue of
A. By Lemma 8.14(4), we see that r(A) = 1. Gershgorin's theorem also
informs us that if A is an eigenvalue of A, then for some i we have
Al IA = 1 a22.
for some subset I of {1,2,. . . ,n}. We are now ready to define the notion
of ideal irreducibility for matrices. In Chapter 9, this definition will be
extended to arbitrary operators; see Definition 9.2.
8.4. Irreducible Matrices 333
Proof. Let A be an irreducible matrix. If A has a zero row, say the then
for each x E we have = 0, i.e., the component of Ax is zero.
This implies that the proper ideal J = {x E = 0} is A-invariant, a
contradictionS
On the other hand, if A has a zero column, say the and we consider
the proper ideal J {x = 0 for all i j}, then A(J) = {0} C J,
which is also impossible.
(ni) factors
From (I + A)x = x + Ax, we see that [(I + A)x]t xt > 0 for each t E I.
To see that [(I + A)x]t = xt + (Ax)t > 0 for some t I, we must show that
(Ax)t > 0 for some t I. To this end, assume by way of contradiction that
(Ax)t = 0 for all t I and consider the non-trivial ideal
J={zECn: zj=0 for all
If z J, clearly Izi < Ax holds for some A > 0, and therefore from
then
lAzi AIzI AAx it follows that = 0 for all t I, i.e., Az E J. In
other words, A(J) J, contrary to the irreducibility of A, and the validity
of our claim follows.
(3) (4) Assume (I + is strongly positive. First, we observe
that A cannot have a zero row. For if a row of A is zero, say the ith row,
then the row of I + A is equal to the basic unit vector But then I + A
8.4.Irreducible Matrices 335
A(I + = (n1)Ak]
=
we thatfor each i) there some integer k (in 1 k n).
such that > 0.
(4) (5) Assume by way of contradiction that there exists some
permutation matrix P such that PtAP where B and D are
square matrices. Then for each k we have
where Bk and Dk are also square matrices having the same sizes as B and
D, respectively. From the above representation of we infer that there is
at least one position (i, j) such that the (i, entry of Ak is zero for each
k. However, the latter contradicts our hypothesis and the validity of (5) is
established.
(5) Assume by way of contradiction that A leaves invariant a
(1)
non-trivial ideal J = {z E = 0 for all i I}. Let I {i1,. . ,
In particular, for each z E J, the column vector QAz also has its last k
components all equal to zero.
Now let us look at the permutation matrix P = the transpose
of Clearly, the last k co'umns of are the column unit vectors
. , In particular, the components of the first m k column vectors
of are all zero at the coordinates So, if the columns of AQt
,
are then the column vectors ci,.. .,Ck all have zero
components at the coordinates ii, . . . , 1, .. , m k i .
the column vector must have (by the choice of Q) its last k components
336 8. Positive Matrices
equal to zero. In other words, the matrix ptAp = QAQt is of the form
A positive;
is positive, i.e., r(A) = )'A > 0.
Proof. (1) and (2) By Corollary 8.20, there exists a strongly positive matrix
B such that AB = BA. Clearly, this matrix AB is also strongly positive.
Indeed, if z > 0, then Bz >> 0, and (since A has no zero rows) ABz>> 0.
Now let Ax AAx for some x > 0 and assume by way of contradiction
that Ax > )t44x. By Corollary 8.21 we have AA > 0. Now consider the
vector y = AAx) and note that y> 0 and Ax = AA(x + y). Hence,
Proof'. (1) By Theorem 8.22, there exists x 0 such that Ax = AAX. This
implies >> Ax for each and so (AA, oc) c UA.
338 8. Positive Matrices
Exercises
1. For a vector subspace J of show that the following statements are
equivalent.
(a) J is an ideal, i.e., J = Jo zJo, where Jo is an ideal in
(b) J is a band, i.e., J = B iB, where B is a band in RTh.
(c) There exists some non-empty subset I of {1, 2, . . . , n} such that
lim
nboo
= fl+OO
lim = Ax > 0,
where the last conclusion holds true by virtue of the strict positivity of A
(Corollary 8.21). This yields = and so Ax ,\0x. By
Corollary 8.23, = = r(A) and this implies limt,t0 = r(A).
(*)
for each k = 1,. . , n. This implies that for each k there exists some angle
0k such that = for j = 1,.. . , m; see Exercise 1 at the end of
this section. But then (*) yields (EL1 Ixj I) = I
(xj and I
so for each fixed k either = 0 for each j or else etk = 1. In either case,
bkj Ibkjf Oforallkandj.
From the definition of B it follows that
B = IBI = A D1 = A,
and therefore akj 0 for all Ic and j. Now taking into account that
AlxI = BJxJ, we see that bkj)fxjI = 0 for all k. Since 0
for each j, the latter implies ak3 = for all k and j, i.e., A = B, and
our conclusion follows. (Another way of proving that A = B is by using
Corollary & 23(3) in conjunction with the inequality 0 <B A.)
from A by deleting its Icth row and kth column. Also, let Pk ( ) denote the
characteristic polynomial of Ak, i.e., det(AI Ak). An easy corn-
putation (using the familiar formula for differentiating a determinant whose
entries are functions) shows that
k=0,...,p1}
is a group, and consequently G = {z E C: = 1}, i.e., G has the desired
form; see Exercise I in Section 7.3. Now invoke Corollary 8.25 once again to
obtain that the whole spectrum o-(A) is rotation invariant under the angle
The simplicity of the roots can be established either directly or by
employing the conclusion of part (6) below.
To see that p can be computed by the described formula, note first
27r
that it follows from Lemma 8.24 that the two matrices A and e are
similar, and so they have the same characteristic polynomial. Computing
the characteristic polynomials yields
PAP') = + + ... + aflk.A
27r(nn1) 21r(nnk)
= + p
+ aflke
2ir(n_n,)
This implies = or e =I for each j. Therefore, p
divides n nj for each j.
To establish that p is the gcd of the numbers n we distinguish two
cases: p = I and p> I. Assume first that p = I, i.e., cTper(A) = {r(A)}. In
this case, if m > I is a divisor of each n then (as above) the matrix
has the same characteristic polynomial as A. This guarantees that
is a point of the peripheral spectrum of A which is different from
a contradiction. Now consider p > I. If m > I is a divisor of each
8. 5. The PerronFrobenius Theorem 343
an eigenvalue by does not change its multiplicity, and our final conclu-
sion follows.
Corollary 8.27. Let A = [aj.:j] be an n x n irreducible positive matrix. If
0 for some i, then crper(A) = {r(A)}.
(*)
g],
where C is an (m x (m 1) upper triangular matrix having along its main
1)
diagonal only eigenvalues of A with moduli strictly less than one. The latter
implies 0 (see Exercise 8 of Section 8.1). Since =B B'
for each m, we see that
At(B_i)t = (B_i)t [1 U]
AB= and
and from this it easily follows that Ac = c and Atbt = Lit. Since A and At are
irreducible and both have spectral radius 1, we infer that all components of
the vectors Li and c are non-zero; see Theorem 8.22(2). Now notice that the
(i, j)th entry of S satisfies = 0 and this guarantees that indeed
> 0 for all i and j.
Corollary 8.30. If A is a primitive matrix with r = r(A), then for each
x > 0 the limit y = exists, is strictly positive, and satisfies
Ay=ry.
Proof. Fix x > 0. By Theorem 8.29, we have Sx = y>> 0.
Moreover, note that Ay = lim = r lirn = ry.
In view of the preceding results one might think that it would be easy
to figure out which finite subsets of complex numbers are the spectra of
irreducible positive matrices. Surprisingly, this is a very difficult problem
and a complete answer is not yet known. Even when we place a bound on
&5. The PerronFrobenius Theorem 345
Exercises
1. Show that m complex numbers Zi,... satisfy zj
=
I
if and only if there exists some angle 9 such that z,, = for each
i=1,...,m.
2. Consider the polynomials:
010 0 0
0 0 1 and 0
100 0 2
0
A
0100.
1
0010
0 0 0
00000010
0 0
00000001
0 0 0 1 0 0
11100000
11010000
10110000
01110000
346 8. Positive Matrices
IQ
[HINT: Note that = . This, in conjunction with
[w;] [1].
Corollary 8.30, imPlies Now use that
100.]
Chapter 9
Irreducible Operators
347
9. Irreducible Operators
operators and their relationships. We shall group these operators into three
categories according to their special properties. The first two categories of
operators will be defined in terms of certain invariant subspacesthe ideals
and bands.
Let T: X X be a bounded operator on a Banach space. Recall that
a vector subspace V of X is invariant under T (or simply T-invariant)
if T(V) V. A vector subspace V of X is said to be hyperinvariant
for T (or simply T-hyperinvariant) whenever V is invariant under every
bounded operator on X that commutes with T, i.e., whenever S E (X)
and ST = TS imply 5(V) V. If X is a Banach lattice and T is a positive
operator, then a vector subspace V of X is called lattice hyperinvariant
for T (or simply whenever V is invariant under every
positive operator on X that commutes with T. As usual, a vector subspace
V of a vector space X is non-trivial if V {O} and V X.
The null ideal NT of a positive operator T: B > F between two Riesz
spaces is the ideal in B defined by
NT={XEB:
T= 0 if and only if NT = E. It should also be obvious that
if B and F are Banach lattices, then the null ideal NT is closed in B.
Moreover, observe that a positive operator T: B F is strictly positive (see
Definition 1.13) if and only if NT = {0}. If T: B F is a positive operator
between two Riesz spaces, then the positive operator T: B/NT F, defined
by T[xJ = Tx, is strictly positive. -
We are now ready to define the classes of ideal and band irreducible
operators. They extend the corresponding notions for matrices; see Defini-
tion 8.16.
Definition 9.2. A bounded operator T: E E on a Banach lattice is:
(1) ideal irreducible if T has no invariant non-trivial closed ideals,
that is, whenever a closed ideal J satisfies T(J) J, then either
J_{O} orJ=E, and
(2) band irreducible if T has no invariant non-trivial bands, i.e.,
whenever a band B satisfies T(B) B, then either B = {O} or
B=E.
The next result indicates that "ideal irreducibility" is a relatively strong
property. In particular, as we shall see, only Banach lattices with quasi-
interior points admit ideal irreducible positive operators.
Theorem 9.3. If T: E E is an ideal irreducible positive operator on a
Banach lattice, then:
(1) T is band irreducible.
(2) T is strictly positive, i.e., x > 0 implies Tx > 0.
(3) P2 has quasi-interior points.
(4) T carries points to quasi-interior points.
Proof. (1) The conclusion follows from the fact that every band is a closed
ideal.
(2) Let T: P2 E be an ideal irreducible positive operator on a Banach
lattice. The null ideal NT of T is closed and T-invariant. Since T is ideal
irreducible, it follows that NT = {0}, which means that T is strictly positive.
(3) By Theorem 4.21 there exists some u > 0 such that
This implies T is ideal irreducible
and {0}, we see that = E. So, u is a quasi-interior point.
(4) Let u> 0 be a quasi-interior point of P2 and let JT denote the closed
ideal generated by the range of T. That is,
JT={xEE: ay>Osuchthatlxl<Ty}.
By Lemma 9.1, JT is a closed -hyperinvariant ideal for T, and so = P2.
Since u> 0 is a quasi-interior point, the principal ideal generated by u
is norm dense in P2. Clearly,
T(E) c JT.
This implies = JT = E, i.e., Tu is also a quasi-interior point of P2.
350 9. Irreducible Operators
A similar result holds true for band irreducible positive operators; see
Exercise 10 at the end of the section. We now introduce the classes of
"expanding" operators.
There are some other classes of operators that are closely related to the
classes of operators introduced above. These are the classes of Krein and
strong Krein operators that were introduced in [8]. Their precise definition
is as follows.
The next result shows schematically the relationships between the vari-
ous classes of positive operators introduced above.
9.1. Irreducible and Expanding Operators 351
Theorem 9.6. We have the following implication scheme for positive op-
erators on a Banach lattice:
Strongly Krein
V
Krein Strongly Expanding
Expanding
Band Irreducible
V
In general, no other implication is true.
T(xi,x2,...) =
Tx(t) = / sin(st)x(s)ds.
JO
It is easy to verify that T is an expanding compact operator. Since Tx(O) = 0
for each x E C[0, 1], this operator fails to be strongly expanding. Also,
notice that if J = {x e C[0, 1]: x(0) = 0}, then J is a closed
ideal and T(J) J holds. This shows that T is not ideal irreducible; for
more properties see Exercise 8 of Section 9.2.
Example 9.10. An ideal irreducible operator which is not weakly expanding.
Consider the operator T in Example 9.7 restricted to co. As noticed
there, T(co) co. Since CO has order continuous norm, the notions of ideal
irreducibility and band irreducibility coincide, and so T: co * co is ideal
irreducible. The final argument in Example 9.7 shows that T: co + co is
not weakly expanding.
Example 9.11. A weakly expanding operator which is neither expanding
nor ideal irreducible.
Consider first the positive operator 8: co p CO defined by
8(xi, X2,...) = (x2 + X1, + +
It should be immediate that S is weakly expanding but fails to be expanding.
Arguing as in the last part of Example 9.7, we see that S is ideal irreducible.
Moreover, since co has order continuous norm, it follows that S is also order
continuous.
Next, extend S to the Banach lattice c of all convergent sequences by
letting
T(xi, x2,...) = S(xi x2 +
where = and e = (1, 1, 1,. .). The reader can verify that
.
Proof. We shall consider only the case when T is positive and ideal irre-
ducible. The other case is similar.
The proof is by induction on k. Assume first k = 0. Let x > 0 and put
u = Sx > 0. From T(Sx) ASx (i.e., Tu we see that c
By the ideal irreducibility of T, we get = E. That is, u = Sx is a quasi-
interior point, and so S is strongly expanding. Now assume that is
strongly expanding. Since = T(Tc_lS) = and
(by the induction hypothesis) is strictly positive, the above argument
shows that is strongly expanding.
Finally, since S is strongly expanding and Tk is strictly positive, it should
be immediate that is also strongly expanding.
Proof. Note that for each A > r(T) the operator R(A, T) is strictly positive
and satisfies T) AR(A, T); see Lemma 6.23(3).
Exercises
Tf(t) = (*)
forallO<t<landTf(O)=f(O). Showthat:
(a) T defines indeed a positive operator on C[O, 1].
(b) T is one-to-oneand hence strictly positive.
(c) T is a Markov operator.
(d) T is not band irreducibleand hence not ideal irreducible either.
(e) The point spectrum of T is the set
14. Let 1 < p < oo. We consider the operator T of the preceding exercise
defined on [0, oc) via the same formula
for all > 0 and Tf(0)=f(0). Show that (**) defines a one-to-one strictly
s
positive operator from oc) to oo) such that =
15. Prove the converse of Corollary 9.14. That is, show that if T: E E
is a positive operator on a Banach lattice such that T) is strongly
expanding for some A > r(T), then T is ideal irreducible. [HINT: If a -
17. Give examples of two positive operators S and T on atomic and non-
atomic Banach lattices such that T is ideal irreducible, ST 5, and
S is a non-expanding strictly positive operator. (This shows that the
inequality TS AS in Lemma 9.13 cannot be replaced by ST AS.)
Theorem 9.19 (de Pagter [260]). Every mom-zero compact positive operator
which is locally quasimilpotent at a mom-zero positive vector has a mom-trivial
closed -hyperimvariamt ideal.
such that
E U
for each ri. From IAjzI and RT = TR we get for each
ri. The latter implies 1IRI1Th Hence,
1 1
0,
and so > 0. E U for each ri, it follows that 0 E U, contrary to
Since
(*). This contradiction completes the proof of the theorem.
Corollary 9.20 (de [260]). An ideal irreducible compact positive
Pagter
operator on a Banach lattice has a positive spectral radius.
It should be pointed out that the compactness assumption in the above
corollary is essential; see the comments at the end of Section 9.3. We are
now in a position to sharpen the previous result by proving the positivity of
the spectral radii of two commuting positive operators.
Corollary 9.21 (AbramovichAliprantisBurkinshaw [7]).If an ideal ir-
reducible positive operator T: E * E commutes with a compact positive
operatorS: E E, then r(T) > 0 and r(S) > 0.
Proof. Let 5, T: E E be two positive operators on a Banach
lattice. Assume that S is compact, T is ideal irreducible, and ST = TS.
From Theorem 9.3, it follows that ST = TS > 0.
360 9. Irreducible Operators
The next results show that the compactness assumption in Corollary 9.20
can be considerably relaxed.
'\'Ve conclude the section with a general result concerning strongly ex-
panding operators.
362 9. Irreducible Operators
Then r(QS) > 0, i.e., the spectral radius of the operatoT QS is positive.
The last four results are due to the authors and 0. Burkinshaw [7].
Exercises
QT={xEX:
Show that QT is a T-hyperinvariant vector subspace. 2 [HINT: Let x, y
in QT and fix 0. Pick n0 such that and for
2
some applications and historical comments regarding QT see [233, 234].
364 9. Irreducible Operators
1 1
The analogue of Theorems 9.26 and 9.19 for lattice hyperinvariant bands
is as follows.
Theorem 9.31. If T : E E is an order continuous compact quasi-
riilpoterit positive operator on a Bariach lattice, then. there exists a non-trivial
-hyperinvariant band B for T. Moreover, if T2 0, then the band B can
be chosen to satisfy T(B) {O}.
Proof. Ifthe null ideal N = {x E: TIrcI = o} is non-zero, then N is (by
Lemma 9.1) a non4rivial T-hyperinvariant band. So, we can assume that
NT = {O} , and hence T is strictly positive. Let JT denote the closed ideal
generated by the range of T in E, i.e.,
JT={XEE:
By Lemma 9.1, JT is a closed lattice hyperinvariant ideal for T. Since
the operator T: JT JT is strictly positive and quasinilpotent, it follows
from Theorem 9.26 that T, as an operator on JT, has a non-trivial closed
-hyperinvariant ideal J JT. Since JT itself is -hyperinvariant, it fol-
lows that J, considered as an ideal of E, is also -hyperinvariant. Now let
B be the band generated by J in E. By Lemma 9.30, every positive op-
erator commuting with T is order continuous. This implies that B is an
-hyperinvariant band for T. To finish the proof, it suffices to show that
B E. To see this, assume by way of contradiction that B = E. Hence,
for each x > 0 there exists a net in J such that 0 x. By the
order continuity of T, we get I Tx, and by the compactness of T we
see that the increasing net is norm convergent in E. It follows that
0. Since J is J, and thus Tx E J for
all x E E. This implies J = contradicting our assumption J
To establish the second part, assume now that T2 0. In this case, we
consider the operator I': E/N E/N defined by i'[x] = [Tx], where N is
again the null ideal of T. Since T is order continuous, N is a band. We
claim that T> 0. Indeed, if we pick x > 0 such that T2x > 0, then Tx N
and so T[x] > 0. Clearly, T is quasinilpotent and compact. Furthermore,
we claim that T is order continuous.
To see this, let J. 0 in E/N. Replacing } by the net of all finite
infima of the net we can assume that 0 j holds in E. Since E is
Archimedean, there exists a net {y,), } of E satisfying 0 y%, for each
a and each A and 0; see [225, Theorem 22.5, p. 115]. From
J. 0, we see that [y,)j = 0, i.e., = 0 for each A. Therefore, replacing
{ by we can assume without loss of generality that J. 0 in
E. By the order continuity of T, it follows that Tx0, J. 0. So, by the order
continuity of the quotient map x 'p [x] (see Exercise 2 at the end of this
section), we obtain T is order continuous.
Band Irreducibility and the Spectral Radius 367
Proof. Fix some A > r(T) and consider the operator Q = R(A, T)T.
Clearly, Q commutes with 5, is u-order continuous, and, by Corollary
is expanding. Therefore, the operators Q and S satisfy the hypotheses
of Theorem 9.35, and thus r(QS) > 0. Finally, using the inequality
r(QS) r(R(A,T))r(T)r(5), we obtain r(T) > 0 and r(S) >
9.3. Band Irreducibility and the Spectral Radius 369
Exercises
1. Show that the conclusion of Lemma 9.30 is false if the "compactness" of
T is replaced by "ideal irreducibility" of T. That is, give an example of
two commuting positive operators T, 8: E E on a Banach lattice such
that:
(a) T is strictly positive, order continuous, and ideal irreducible.
(b) S is not order continuous.
2. Let J be an ideal in a Riesz space E. Show that J is a band if and only if
the quotient map x [x], from E onto E/J, is order continuous. [HINT:
Assume that J is a band, that .1. 0 in E, and that [x] in E/J
for all a and some x E E+. Now use the fact that the quotient map
is a lattice homomorphism (see Exercise 2 of Section 3.4) to obtain that
0
3. Prove the following analogue of Corollary 9.28: Let T: E E be a a-
order continuous positive operator on a Banach lattice. If the commutant
of T contains a cr-order continuous band irreducible positive operator and
also a cr-order continuous compact positive operator, then T(T) > 0.
370 - 9. Irreducible Operators
and put
(1) Let A = be a real n x n matrix such that > 0 for all (i,j).
Then A is a strong Krein operator from to
(2) Assume that C(11) admits a strictly positive linear functional f.
(This is equivalent to assuming that there exists a regular Borel
measure whose support is If u is a unit, then the rank-one
operator T = f u is a strong Krein operator.
More generally, any finite-rank operator T =
where each is a strictly positive linear functional on C(11), uj > 0
for each i, and is a unit, is a strong Krein operator.
Tf(t) =
Jci
Two useful properties of Krein operators are included in the next lemma.
Lemma 9.40. If T: p C(11) is a Krein operator, then:
(2) TTh is a Krein operator for each n 1 (and so TTh Al for each
n 1 and all scalars A 0).
Proof (1) Let a unit. Since T is a Krein operator, there exists some
be
positive integer k 1 such that v = is also a unit. Next, choose some
> 0 such that $ 'yu and note that v = 7Th holds. This
implies that T'u is a unit.
(2) Using (1) and an inductive argument, we see that carries units
to units for each k. Now fix n 1 and let x> 0. Choose some k such that
is a unit. But then (TTh)kx = is a unit, which proves that
is a Krein operator. El
Notice that from (1) above, it follows that if T is a Krein operator and we
let v = TI, then the operator x is a Markov operator. Regarding
the spectral radius of a Krein operator, we have the following.
Theorem 9.41. If T: p is a Krein operator, then:
(1) T has a positive spectral radius, i.e., r(T) > 0.
1
(2) r(T) = lim for each x> 0.
Proof. (1) By Lemma 9.40, Ti is a unit. This implies r(T) > 0; see
Exercise 2 at the end of this section.
(2) Let x > 0. Fix some k such that 'ii = TJcx is a unit. This implies
that the formula = inf{A 0: AtL} defines an equivalent norm
Krein Operators on -spaces 373
1
Since and are equivalent norms, we get n = r(T)
and the proof is finished.
Proof. (1) Assume that Tx0 = for some xO > 0. Since T is a Krein
A0x0
operator, there exists some positive integer Ic such that is a unit. From
Tkxo = (Ao)kxo, it follows that A0 > 0 and that xO is a unit.
Now from Theorem 9.41, we obtain
1 1 1
r(T) = lim
fl-400
= lim = A0 lim
TL400
= A0.
Also, consider the continuous mapping F: (Ca, w*) (Ca, w*) defined by
Then:
A proof of the above result that does not use a fixed point argument can
be found in [185]. The next result generalizes the previous corollary to a
family of commuting operators.
Theorem 9.47 (M. G. Krein). Let be a family of pairwise com-
muting positive operators on some C(11)-space. Then the family of
a common positive eigenvector. That is, there exist
o < and a family of non-negative scalars such that
T* "F T* ( crrL+191 \
A1________
'ar'a2
.
*
for each x E C(12).
= E E A, xj with x=
m m m
Tm = ... .. .
k1=dk2=l
i=l k1=l
it m m m m
= ... fJ -
i=1 k1=l
flrpkj(rpm+1 m
U
=
Consequently, from the triangle inequality, we infer that
um rn
Since e is assumed to be a unit, there exists some 0 such that u
This implies
1= = IITm(U)tIu <T/IITrn(e)IIu
for each rn, which is impossible. Therefore, Y fl U = 0.
Since U is open, the Separation Theorem guarantees the existence of
some non-zero b E satisfying = 0 for all y E Y and 0 for
all x e U. Taking into account that U = we conclude that q5> 0.
From x E Y, it follows that
(x, = (Tax x, 0,
for each x e Thus, = b for each a E A, and this shows that
0 is a fixed point for the family of operators
Corollary 9.49. Let T: be a positive operator. If some A > 0
is an eigenval'ae of T having an eigenvector which is a 'anit, then the adjoint
operator T* also has A as an eigenval'ae with a positive eigenvector.
9. Irreducible Operators
If all operators in the previous result are contractions, then we have the
following companion to Theorem 9.48.
Theorem 9.50. Let {Ta} be a family of pairwise commuting positive con-
tractions on C(cfl. If there exists a non-zero vector v e such that
v for each a, then there exists a non-zero positive linear functional
e satisfying = for each a.
Proof. Start by fixing some non-zero vector v E C(12) such that =v
for each a. Next note that if k1, k2,.. , are arbitrary natural numbers
and al, a2,. . . , are arbitrary indices, then
. lxii,
where denotes the sup norm.
Now we consider the same vector subspace Y of as in the proof of
Theorem 9.48. Assume that e E Y fl U, and then for each m construct the
operator Tm. If we let c = rnax{ Jjxj!i: i = 1,.. , n}, then it follows that
Now if > 0 satisfies IvI then
for each m, which is impossible. The rest of the proof is the same as the
proof of Theorem 9.48.
Corollary 9.51. If a positive contraction T: has 1 as an
eigenvalue, then there exists some 0 <4 e such that T*q5 =
Exercises
1. Let T: be a positive operator. Show that for each unit
uE we have
1
= and r(T) =
Invariant Subspaces
One of the most famous and still unsettled problems in functional analysis
and modern mathematics is the so-called invariant subspace problem
that is described by the following simple question:
o When does a bounded operator on a Banach space have a non-trivial
closed invariant subspace?
To recall some basic definitions related to the above question, let us
suppose that T: X X is a bounded operator on a Banach space and V is a
subspace of X.' Then V is non-trivial if V {O} and V X. We say that
V is invariant under T (or simply T-invariant) if T(V) V. Also, V is
said to be hyperinvariant for T (or simply T-hyperinvariant) whenever
V is invariant under every bounded operator on X that commutes with T,
i.e., if S e (X) and ST = TS imply that 5(V) V. If X is a Banach
lattice and T is a positive operator, then V is called lattice hyperinvariant
for T (or simply -hyperinvariant) if V is invariant under every positive
operator on X that commutes with T.
It is very easy to show (see the beginning of the next section) that for
non-separable Banach spaces the invariant subspace problem has an imme-
diate affirmative answer. For finite dimensional complex Banach spaces the
invariant subspace problem has an affirmative answer too. To see this, let
T: X X be a bounded operator on a finite dimensional complex Banach
space (of dimension greater than one). There is nothing to prove if T is a
multiple of the identity operator, since then each subspace is invariant. So,
1
Unless otherwise stated, throughout this chapter X will denote an arbitrary real or complex
Banach space of dimension greater than one. Similarly, E will denote a real Banach lattice of
dimension greater than one.
381
382 - I Invariant Subspaces
The next two results deal with the existence of invariant subspaces for
operators on finite dimensional vector spaces. As we shall see, the situation
is different between the real and complex cases.
Corollary 10.6. Every non-scalar operator on a finite dimensional complex
Banach space has a non-trivial hyperinvariant subspace.
Proof. Let T: X X be a non-scalar operator on a finite dimensional
complex Banach space, and consequently T has an eigenvalue A e C. Using
that T is a non-scalar operator, we infer that the eigenspace NA is non-trivial.
By Lemma 10.5, this eigenspace NA is T-hyperinvariant.
10.1. A Smorgasbord of Invariant Subspaces 385
X\{0} and, in particular, they cover the compact set K(U0)._So, we can find
a finite number of operators A1,... ,Ak E A such that K(U0) c
Let c = and note that c> 0.
Since Kx0 E K(Uo), there exists some 1 ji k such that Kx0 E
and thus x1 = Kx0 E U0. Then Kx1 K(Uo) and so there exists some
I k such that x2 = E (J0. Proceeding inductively, we can
select a sequence of indices in {I, 2,.. k} such that the sequence
. ,
Exercises
1. (Schaefer [287]) Show that every linear operator on an infinite dimen-
sional vector space has a non-trivial invariant subspace. [HINT: Let
T: X X be a linear operator on an infinite dimensional vector space.
Fix some x 0 and consider the T-orbit space (9T of x. If (9T
then (9T is a non-trivial T-invariant subspace. So, assume that = X.
In this case, it follows that the set of vectors {x, Tx, T2x,. . is linearly
independent. This means that every vector y E X has a unique repre-
sentation of the form y = witl1 the )'tj equa' to zero for all
but finitely many i. Now define the non-zero linear functional on X by
q5
S(x1,x27x3,x4) =
then S commutes with T and leaves neither nor Vr invariant. On
the other hand, if a is a non-zero angle such that + is an irrational
multiple of 2ir, then the operator R: 1R4 defined by
R(xi,x2,x37x4) =
commutes with T and is not R-invariant for all 'y > 0.1
3. Let T: X X be a bounded operator on a Banach space. Denote
by Lat(T) the set of all closed T-invariant subspaces of X. Show that,
under the standard set inclusion. Lat(T) is an order complete lattice with
smallest and largest elements.
392 10. Invariant Subspaces
Since (in view of the compactness of K) the set K(Uo) is compact, there
exist operators A1,... , A such that
IIAiy-xo(1<1}.
Since x E Uo and (1(x) > 0 imply E Uo, it follows from the convexity
of U0 that Uo. In addition, because is a convex combination
of the vectors A1Kx,.. and E we see that
belongs to C = But is a norm totally
2jf X is a complex Banach space then when taking the closed convex hull of C, we consider
X as a real vector space.
10.2. The Lomonosov Invariant Subspace Theorem 395
And now we are ready to state and prove the two famous Lomonosov
invariant subspace theorems.
Theorem 10.19 (Lomonosov [211]). If a non-scalar operator T on a com-
plex Banach space commutes with a non-zero compact operator, then T has
a non-trivial closed hyperinvariant subspace.
Proof. Assume that T E (X) a non-scalar operator and that for some
is
non-zero compact operator K E (X) we have TK = KT. To establish the
existence of a non-trivial closed T-hyperinvariant subspace, according to
Lemma 10.14, it suffices to show that the corninutant of T is non4ransitive.
To verify this, assume by way of contradiction that the commutant {T}"
is a transitive algebra. So, by Theorem 10.18, there exists some A E {T}"
such that the operator AK has a non-zero fixed point. Let
F={xEX: AKx=x}.
Clearly, F is a non-zero closed AK-hyperinvariant subspace. Since AK is
a compact operator and it is the identity on F, it follows that F is finite
dimensional. Since T commutes with AK and F is AK-hyperinvariant, we
see that F is So, from T(F) F and the fact that F is a finite
dimensional complex Banach space, it follows that the operator T: F F
has an eigenvalue C.3 Now let
NA={XEX:
and note that NA is a non-zero closed {T}"-invariant subspace which is dif-
ferent from X since T is a non-scalar operator. However, this contradicts
our assumption and the proof of the theorem is complete.
This is the place where we use that X is a complex Banach space. When X is a real
Banach space, there is no guarantee that the operator T: F + F has a real eigenvalue.
396 -- 10. Invariant Subspaces
Exercises
1. Give an example of a polynomially compact operator which is not com-
pact.
2. If is a compact Hausdorif space without isolated points and a function
E C(Q) is one-to-one, then show that the multiplication operator
does not commute with any non-zero weakly compact operator.
3. Show that every multiplication operator on a is a Lomonosov
operator.
4. Assume that (cl, >, is a cr-finite measure space such that there exist
two disjoint sets A, B E with 0 <cc and 0 <cc. Show
that every multiplication operator on an with 1 p oo is
a Lomonosov operator.
5. Show that every Markov operator T on an AM-space with unit commutes
with a rank-one positive operatorand hence T is a Lomonosov operator.
6. Consider the operator T: C[1, 1] + C[1, 1] defined by Tx(t) = x(t)
for each x E C[1, 1] and all t E [1,1].
(a) Show that T is a non-scalar surjective lattice isometry which is also
a Markov operatorand so, by Exercise 5 above, T commutes with
a rank-one positive operator and is a Lomonosov operator.
(b) Exhibit a rank-one operator on C[1, 1] that commutes with T.
(c) Show that the compact operator 8: C[ 1, 1] + C[ 1, 1], defined by
The results in this section (as well as the results in the subsequent sections)
will provide a considerable amount of evidence to expect the affirmative an-
swer to this conjecture. For a while it was thought that the modulus of some
of the quasinilpotent operators on (the real) 4 without non-trivial closed
invariant subspaces constructed by Read [275, 276, 277] might provide
a counterexample to this conjecture. However, V. G. Tftoitsky [319] has
shown (see Corollary below) that the modulus of any quasinilpotent
Read operator on has an eigenvalue, and so its modulus necessarily has
a non-trivial closed invariant subspace.
[A,B] =AB-BA.
jByJ BJyJ
see Lemma 4.15. In particular, for each x 0 there is some m such that
lixo - xO A <1.
Since the function z Fk xO A riA(lzI) is continuous, we see that the set
{z E E: <i} is open for each ri. In view of 0
the above arguments guarantee that
k(Uo)cU{zEE: <1}.
Since the sets {z E E: IxoxoAriA(lzt)1l < i} are increasing as ri increases,
the compactness implies that
K(Uo) {z E E: lIxo x0 A rnA(lzl)11 <i}
for some m. In other words, there exists some fixed rn such that x E K(Uo)
implies that xO A rnA(lxI) E Uo.
In particular, we have x1 = E Uo. So, Kx1 E K(Uo) and
this implies x2 = xO A E Proceeding inductively, we obtain a
sequence of positive vectors in U0 defined by xlHi = xO A
Now we claim that the inequality
0
holds for each ri. The proof is by induction. For ri = the desired inequality
1
Moreover, condition (2) can also be relaxed. The proof of the next result is
very similar to that of Theorem 10.24 and is omitted.
Theorem 10.25. Let B: E E be a positive operator om a Bamach lattice
amd assume that there exists a positive operator 5: E + E such that:
(1)
1
(2) There exists some xO > 0 satisfyimg liminf = 0.
11-4(X)
(3) S domimates a mom-zero AM-compact operator.
Them the operator B has a mom-trivial closed imvariamt ideal.
J={yEE: A>Osuchthatlyl<AAx0}.
Clearly, J is a (B + 8)-invariant ideal. Since 0 B, S S + B, it
follows that this B and S.
ideal is invariant under both
If S is not strictly positive. then consider the null ideal N8. It is a non-
trivial closed S-invariant ideal in E. Since S e (B], we know that B E [5)
and so Ng is also B-invariant: see Lemma 10.23.
Consequently, we can assume that S is strictly positive. Therefore, for
any -y e F the operator K is non-zero, where R,. denotes the stan-
dard rank-one projection onto the one-dimensional subspace spanned by
Clearly, K is compact and 0 <K < 5. Now apply Theorem 10.24.
B is compact, whence
HBIThM ess = lBtm + <
IIBThII.
Consequently,
We shall close the section with two interesting cases where the existence
of an invariant subspace of a positive operator can be obtained without any
compactness assumptions.
Theorem 10.34. If is a compact Hausdorff space, then every non-s calar
positive operator on has a non-trivial closed hyperinvariant subspace.
In particular, each Krein operator has a non-trivial closed hyperinvari ant
subspace.
be the corresponding powers of B for these points, i.e., flj = Now con-
sider the operator
S= + + Bn'k, (**)
and note that for each t E F we have
= k
[Sxo] (t) (t) min{ct17. . .
> 0.
Exercises
1. Let S: E E be a positive operator on an arbitrary Banacli lattice such
that Sxo holds for some vector x0 > 0 and some scalar 'y > 0.
Show that S is not locally quasinilpotent at
2. Let J be an ideal in a Banacli lattice E. If J is invariant under some
positive operator B: E + E, then J is also invariant under every operator
dominated by B.
3. Give an example of a pair of non-commuting positive operators A and B
such that A E [B), that is, AB BA> 0.
4. For a positive operator B: E + E on a Banach lattice establish the
following.
(a) The range ideal RB of B (i.e., tile ideal generated by the range of
the operator B) is given by
RB = {y E E: 2 x E such that Bx }.
(b) The range ideal RB is -hyperinvariant.
(c) For each A E [B) the range ideal RA is B-invariant.
(d) The null ideal NB is A-invariant for each A E [B). In particular,
NB is L?hyperinvariant.
(e) Present an exaniple of an operator A E (B] for whiicli tue ideal NB
fails to be A-invariant.
(f) If B is and [B) contains an ideal irreducible operator, then
B is strictly positive.
5. Assume that K: E + E is a compact operator on a Banach lattice suchi
that Kx0 0 for some x0 E E. Show that there exists a compact operator
C such that Cx0 > 0 and for each x E E. [HINT: Consider
the range ideal RK of K. Using the compactness of K, establish that the
norm ciosure of RK iias a quasi-interior point. Tiien apply Lemma 4.16.]
6. For a positive operator T: E + E on a Banacii lattice let
QtT = {A E (E): B such that AxJ BtxI for all x E
remains true, and so, in view of the compactness of K(U0 fl [0, xo]), there
exists some m such that
K(Uon {z E: X0AmA(IzI)II < 1}.
Now the sequence defined recursively by = and
xn+1 = A n 1, 2, . . .
satisfies E (Jon [0, x0] for each n. The rest of the proof remains basically
the same.]
Prove Theorem 10.26.
10. Recall that a bounded operator T: X Y between Banach spaces is a
DunfordPettis operator if 0 in X implies 0 in Y.
Show that every positive DunfordPettis operator on a Banach lattice
which is quasinilpotent at a non-zero positive vector has a non-trivial
closed invariant ideal. [HINT: Let B: E E be a positive Dujiford
Pettis operator on a Banach lattice which is quasinilpotent at a non-zero
positive vector. We can assume that B is strictly positiveotherwise
NB is a non-trivial closed invariant ideal. Now note that B carries order
intervals onto relatively weakly compact sets [30, Theorem 19.12, p. 339],
and so B2 is a non-zero AM-compact operator.
and hope that whenever this notation is used it will not cause any confusion
with the standard Cartesian product notation. The commutant of C is the
unital algebra of operators defined by
C'={Aer(X): AC=CA for all CEC}.
A subspace V of X is said to be C-invariant if V is C-invariant for each
C E C. A collection C of operators is said to be non-transitive if there
exists a non-trivial closed C-invariant subspace. Otherwise, the family C is
called transitive.
We now extend the concept of local quasinilpotence to collections of
operators.
Definition 10.37. A family C of operators in (X) is said to be:
(1) (locally) quasinilpotent at a point x E X if lim 0
and
(2) finitely quasinilpotent at a point x E X if every finite subcol-
lection of C is (locally) quasinilpotent at x.
Qr={xEX:
It was shown in Exercise 5 of Section 9.2 that QT is a T-hyperinvariant
subspace of X. To generalize this to a collection of operators C, we let
= {x E X: C is finitely quasinilpotent at x}.
As in the aforementioned exercise, we can prove the following result.
Lemma 10.38. The set is a vector snbspace of X and this subspace is
both C-invariant and C'-invariant.
UI 4. Invariant Subspaces of Families of Positive Operators -- 411
For the rest of the discussion in this section we shall assume that:
C denotes a non-empty collection of positive operators on a Banach
lattice E.
The presence of the order structure on E leads naturally to a modification
of the set Namely, we let
a finite subset of C.
If we let = {C1, , C}, then
. . . = {C1, . . , C
n E N. Therefore, from
1
n+1
< =
412 10. Invariant Subspaces
we get Cx f
0. This shows that Cx E and hence the ideal
is C-invariant.
Now note that for each 1 k. This implies at once that
i
for each operator A E we have AT TA, and therefore
<
D1D2 T1,kSl,k] =
k=1 i=1 k=i i=1
Since [C), it follows that (see Exercise 5 at the end of the
section) and hence Therefore,
D1D2
k=i
10.4. Invariant Subspaces of Families of Positive Operators 413
Recall that for x E E the symbol DcX denotes the orbit of x under the
action of that is, D E VC}. To continue our discussion, we
shall introduce one more notation.
. For each x E E we will denote by [DC x] the ideal generated by DCX.
As we shall see, the ideals [Dcx] will provide a source of invariant subspaces.
Lemma 10.42. Each ideal [DCX] is both C-invariant and [C)-invariant.
ICyL
If
<
K(F)c U {yEE:
AEVc
All sets in this union are obviously open, and so we can find a finite subcover.
That is, there exist operators A1,.. , E such that
<
Since x0 E Lemma 10.43 guarantees that the operator A E is
quasinilpotent at x0, and so 0. In particu-
lar, 0. However, the latter conclusion contradicts that C U0
and 0 U0 = U0, and the proof of the theorem is complete.
= sup
-hyperinvariant for B.
Some special cases of Corolaries 10.45, 10.47, and 10.48 have been ob-
tamed by M. Jahandideh [158]. Our next goal is to extend Theorem 10.27
to collections of positive operators. This requires some preliminary work.
Recall that a subset 5o of a multiplicative semigroup S is said to be a
(two-sided algebraic) semigroup ideal in S if for each T E So and for
each S E S the operators TS and ST belong to Clearly, any algebraic
semigroup ideal is a multiplicative semigroup in its own right. Since the
intersection of any family of algebraic semigroup ideals is again a semigroup
ideal, it follows that for any collection of operators C there exists a smallest
semigroup ideal containing this subsetthis semigroup is called
the semigroup ideal generated by C.
The following simple result provides a well-known method of proving
the existence of invariant subspaces for semigroups of operators; see [271,
Lemma 1] and [114, Lemma 4.6].
Lemma 10.49. Let S be a multiplicative group of continuous operators on
a Banach space and let 5o be a non-zero semigroup ideal in S. If So has
a (common) non-trivial closed invariant subspace, then S also has a non-
trivial closed invariant subspace.
Moreover, ifS consists of positive operators on a Banach lattice and the
semigroup ideal S
has a closed invariant ideal.
D }.
We will also need the multiplicative semigroup SCU[c) generated by the dollec-
tionC U [C). We claim that Indeed, take an arbitrary operator
A E 8cu[c) and recall that the elements of the semigroup SCU[C) are just the
finite products of the operators in C U [C). From the definition of [C) we
also know that for any operators C E C and T E [C) we have CT TC.
This immediately implies that the operator A is dominated by an operator
of the form T1 .. TmCi . Ck for some T1,... ,Tm E [C) and C1,.
. Ck E C.
. ,
us verify that
Let contains some operator that dominates a non-zero
compact operator. As we know (see Exercise 5 in Section 10.3), there exists a
compact operator K1 such that IKixl for each x E and K1x0 > 0.
Since = {0}, we can find some C E C satisfying CK1x0 > 0. Hence,
CK1 is a non-zero compact operator on E. Moreover, for each x E E we
have < CIKxI CT0 IxI, establishing our claim since
CT0 E
We have shown by now that the collection of positive operators is
finitely quasinilpotent at x0 > 0 and contains some operator dominating
a non-zero compact operator. That is, satisfies all the hypotheses of
Theorem 10.44. Therefore, according to this theorem, has a non-trivial
closed invariant ideal. But 5o is a semigroup ideal in 5Cu[C)' and so from
Lemma 10.49 we see that 5cu[C) also has a non-trivial closed invariant ideal.
Recalling that C U [C) c 8Cu[C)' we infer that C and [C) have a common
closed invariant ideal. This completes the proof.
j = {x xJ where S and
}
(*)
Proof. For simplicity let T = Assume first that there exists some
> 0 such that w(t)J for each t E I'. So, the function t ' is
Now notice that the sequence lies in the closed unit ball of C(F)*,
which is w*_compact. So, there exists a subnet } of such that
for some measure This and (*) imply at once that
r(T) = n-+oo
lim lim
a
We will establish next that the measure ji is the normalized Lebesgue
measure on I'. To this end consider the operator S: C(F) * C(T) defined
by Sx(t) = x(tz). From (x, = =
it follows that IIS*/#Ln This implies IS*,an in
particular we get 0. From ji, we get
and so = The latter is equivalent to
Jr/
Sxd/i= /xd/i
Jr
for each x E C(F).
Using (**) we can easily show that the measure is rotation invariant.
To this end fix an arbitrary s E I'. Since the argument of z = is an
irrational multiple of 27r, there exists an increasing sequence of natural
numbers such that s. It follows that x(st),aud-
so from (**) and from the Lebesgue Dominated Convergence Theorem we
get
For the general case, consider the sequence of continuous functions {wk}
given by wk = wi + and let Tk = Twk. From for all n we
get r(T) r(Tk) = efr IWkkiIi
for each k. Letting k oo yields
r'(T) (tt)
fin [Iw(t)w(tz) . . .
(]
=
ln I
'Jr
r(T)
Combining this with (if), we see that the formula r(T) = efr In wi d,i is true
in this case too, and the proof is finished.
r(Tw) = sup
(a) O<wo(t)<lforeachtEF.
(b) t 1 is the oniy point at which wO vanishes.
Exercises
1. Let S be a multiplicative semigroup of operators on a Banach space and
let S U {I}, where I is the identity operator. Show that
S and have the same in-
variant subspaces and therefore, when dealing with invariant subspaces
of a multiplicative semigroup, we can always assume that the semigroup
is unital.)
2. Let S be an additive sernigroup of operators on a Banach space and let
= S U{0}, where 0 is the zero operator. Show that is again an addi-
tive semigroup. (Clearly, S and have the same invariant subspaces and
therefore, when dealing with invariant subspaces of an additive semigroup,
we can always assume that the seniigroup contains the zero operator.)
3. Prove Lemma 10.38.
4. Let C be a collection of positive operators on a Banach lattice E and let
= flCECNC. Show that is invariant under C and [C).
5. Let C be a collection of positive operators on a Banach lattice and let
be the multiplicative semigroup generated by C. Establish the following
identities: [C) = [Se), (C] = (Se], and C' =
6. Let C be a commutative collection of positive operators on a Banach
lattice. If C is quasinilpotent at some vector > 0, then show that C is
finitely quasinilpotent at
7. Finish the details of the proof of Theorem 10.51.
10.5. Compact-friendly Operators 425
r(C) = inf
nEN
and
Proof. Without loss of generality we can assume that B < 1. Pick ar-
bitrary scalars > 0 that are small enough so that the positive operator
T= exists and lB + TIl <1. Since [B) is a norm closed additive
semigroup, it follows that T E [B). The same arguments show that for each
n E N the operator (B + T)Th also belongs to [B) and, consequently, -the
positive operator A = + T)Th belongs to [B) too.
For each x > 0 we denote by the principal ideal generated by Ax,
that is,
Jx{YEE: ly!AAxforsomeA>0}.
The obvious inequality x Ax implies that x E and so this ideal is
non-zero.
Observe next that the ideal is (B + T)-invariant. Indeed, if y E
then AAx for some 0 and so we have
00 00
for each x > 0, that is, Ax is a quasi-interior point in E for each x > 0. (In
the terminology of Chapter 9, A is a strongly expanding operator.)
Fix three non-zero operators R, K, C: E E with K, R positive, K
compact, and satisfying
RARARIxI
for each x E E.
Let S RARAR. We will verify that the operator S e [B). To see this
recall that A [B) and that the operator R also belongs to [B) because
it commutes with B. Since [B) is a multiplicative semigroup, the operator
S = RARAR necessarily belongs to [B).
428 " 10. Invariant Subspaces
for each x > 0. Our next step is to show that the assumption (*) implies
that [B) contains an operator that dominates a non-zero compact operator.
Fix three non-zero operators R,K,C: E * E with K,R positive, K
compact, and satisfying
Since C 0, there exists some > 0 such that Cx1 0. This means
that at least one of the vectors (Cxi)+ or (Cxi)_ is non-zero, and therefore,
since E is Dedekind complete, there exists an operator M1 dominated by
the identity operator such that x2 = M1Cx1 > 0. Put = M1C and note
that the operator ii is dominated by the compact positive operator K and
by the operator R.
Since is norm dense in E and C is non-zero, there exists some y e
and an operator A1 E [B) such that 0 < y A1x2 and Cy 5L 0. Using again
the hypothesis that E is Dedekind complete, we can find operators M and
M2 dominated by the identity operator and such that y = MA1x2 and
X3 = M2Cy = IvI2CMA1x2 > 0. Let 112 = and note that the
operator 112 is dominated by the compact operator KA1 and by the operator
RA1.
If we repeat the preceding arguments with the vector x2 replaced by X3,
then we obtain an operator A2 E [B) and an operator 113: E f E which
satisfies 113X3 > 0 and which is dominated by the compact operator KA2
and by the operator RA2.
Consider the operator 113112111. It is non-zero as 113112 = 113X3 > 0.
Moreover, as shown above, each operator (i = 1,2,3) is dominated by a
compact operator and therefore Theorem 2.34 guarantees that the operator
113112111 is necessarily compact. A straightforward verification also shows
that
113112111x1 RA2RA1R(IxI)
for each x E E.
Finally, consider the operator S = RA2RA1R. We claim that S e [B).
To see this, recall that A1, A2 e [B) and also R E [B) because R commutes
with B. Since [B) is a multiplicative semigroup, the operator S = RA2RA1R
necessarily belongs to [B).
It remains to notice that the collection C = {B} consisting of a single
operator B satisfies the conditions of Theorem 10.50. Indeed, C is quasinil-
potent at > 0 and [C) = [B) contains the operator S that dominates a
non-zero compact operator. By this theorem there exists a non-trivial closed
ideal that is invariant under [B). This completes the proof.
10. Invariant Subspaces
a a3 ai.
10.5. Operators 431
T. And [34] and H. J. Krieger [194] proved that each positive irreduc-
ible integral operator on an has a positive spectral radius. In the
"invariant subspace" terminology this means that every positive quasinil-
potent integral operator on an Lu-space has a non-trivial closed invariant
subspace; see Corollary 9.37. This was the first result on the invariant
space problem in the framework of Banach lattices. It should be clear that
Corollary 10.61 is a substantial improvement of the AndKrieger theorem.
For Banach function spaces CorollarylO.61 specializes to the following result.
Corollary 10.62. Let E be a Banach function space associated with a
finite measure space ,i) and let 5: E E be a positive integral oper-
ator with kernel S(.,.). If S is quasinilpotent at some positive vector, then
432 1 0. Invariant Subspaces
Tx(t) = xE E,
where w(.,.) is an arbitrary bounded x ,a-measurable function, has a com-
mon non-trivial closed invariant ideal.
J
e + (yj.
(2) Assume that is a disjoint sequence and let {fTh} be a subse-
quence of By part (1) there exists a subsequence {gm} of such
10.5. Compact-friendly Operators 433
A*x*1 and I
K*1x*1
for each E Exercise 2 in Section 2.4. Now the following three
see
properties can be verified in a rather straightforward way.
(1) For each the support of the measure is contained in the
Exercises
1. Prove that the following operators are all compact-friendly:
(a) compact positive operators,
(b) positive operators commuting with non-zero positive compact
ators,
(c) positive operators that dominate non-zero compact positive opera-
tors,
(d) positive E*_integral operators (in particular, abstract integral opera-
tors), where E is a Dedekind complete Banach lattice,
(e) positive integral operators.
2. Let B: E * E be a positive operator on a Banach lattice. If there exists
a non-trivial Binvariant projection band L\ in E such that Bx = 0 for
each x E L\d, then show that B is compact-friendly.
3. (Sirotkin) For a non-zero positive operator T: E E on a Banach lattice
establish the following.
(a) If K E (E) is a non-zero compact positive operator, then T + K is
compact-friendly.
(b) There exists a operator B E (E) such that T B.
(c) If E is Dedekind complete, then there exists a compact-friendly op-
erator S such that 0 < S < T.
4. Let be a positive one-to-one function in C(c1), where is a compact
Hausdorif space without isolated points. Show directly, i.e., without us-
ing Theorem 10.65, that the multiplication operator Mc1, is not compact-
friendly.
5. A positive operator B: E E on a Dedekind complete Banacli lattice
is called a generalized Harris operator if some power of B is not
disjoint from the band (E* of operators on E, i.e.,
if some power of B dominates a non-zero positive operator.
Recall that if some power of B is not disjoint from the band of
abstract integral operators, then B is called a Harris operator. Of course,
each Harris operator is a generalized Harris operator.
For a generalized Harris operator B: E E establish the following.
(a) B is compact-friendly.
(b) If B is quasinilpotent at a non-zero positive vector, then B has a
non-trivial closed ideal.
10.5. Compact-friendly Operators 435
6. Show that Theorems 10.55 and 10.60 remain true if we replace the as-
1 1
sumption lim = 0 with limiaf =
7. Let 4 p < co) be a continuous operator with modulus.
(1
Assume that there exists a non-zero positive operator 5: 4 4 such
that:
(a) S commutes with the modulus of T.
(b) S is quasinilpotent at a non-zero positive vector.
Show directly, i.e., without using Theorem 10.24, that T has a non-trivial
closed invariant ideal.
8. (Sirotkin) Let B: E + E be a compact-friendly operator on a Banach
lattice such that no positive operator commuting with B dominates a
non-zero compact operator. Show that for each at most countable family
of positive operators commuting with B there exists a non-trivial closed
ideal that is B-invariant and ,F-invariant.
9. [319]) Show that every essentially quasinilpotent compact-
friendly operator on a Banach lattice has a non-trivial closed invariant
subspace.
10. (Troitsky [319}) Let 5: E E be a quasinilpotent integral operator on
a Banach function space such that S is compact. Then each of the
operators Sj and 5+ has a non-trivial closed invariant subspace.
11. Show that the set of all compact-friendly operators on a Banach lattice
E need not be closed or open in (E).
12. This exercise presents an example of a HubertSchmidt operator on
L2 [0, 1] which is one-to-one, positive, locally quasinilpotent at a posi-
tive vector, but fails to be quasinilpotent. The construction of such an
operator will be done in three steps as follows.
(a) Let 0 a <b < 1 and 0 c 1 [see Figure 1(a)] and consider
the integral operator T: L2 [a, bJ r L2 [c, d] defined by
b
Tx(t)=f sin(st)x(s)ds,
b
Tx(t)=f sin(st)x(s)ds, a<t<b,
value K(s, t) is equal to sin(st) if (s, t) lies in each one of the shaded
rectangles shown in Figure 1(b) and is zero everywhere else. The
436 10. Invariant Subspaces
t
1
t a1,
a2
a3
1
a b
IT '11'
a3 a2 al
(a) (b)
C = = : o for each n = 1, 2, .
. .}.
We are ready to present a version of Theorem 10.24 that holds true for
positive operators on a Banach space with a basis.
Theorem 10.66. Let X be a Banach space with a basis and let T: X X
be a continuous positive operator. If T commutes with a non-zero positive
operator that is quasinilpotent at a non-zero positive vector, then T has a
non-trivial closed invariant subs pace.
PTmAXk = 0 (*)
for each m 0. To see this, fix m 0 and let PTmAXk = axk for some non-
negative scalar a 0. Since P is a positive operator and the composition
of positive operators is a positive operator, it follows that
0 aThxk = (PTmA)nXk = TtmThAThxk TtmThAThy0.
For the next result we need to recall the notions of the strong operator
topology on (X, Y). The strong operator topology on (X, Y) is the
Hausdorff locally convex topology 'r8 generated by the family of seminorms
{Px}XEX defined by = JJTxJJ for each T e (X, Y). Notice that a net
{TA} (X, Y) satisfies TA T in (X, Y) if and only if JJTAx k 0
for each x E X.
One can add arbitrary weights to the matrix representing a quasinil-
potent positive operator and still be guaranteed that a non-trivial closed
invariant subspace exists.
Theorem 10.68. Let X be a Banach space with a basis. Assume that a
positive matrix A = defines a continuous operator on X that is quasi-
nilpo tent at a non-zero positive vector. If for a double sequence } of
scalars, the weighted matrix defines a continuous operator B on X,
then the operator B has a non-trivial closed invariant subs pace.
10. 6. Positive Operators on Banach Spaces withBases 439
=0. (**)
for all y E Y. The latter shows that Y is a non-trivial closed vector subspace
of X that is invariant under each operator R: X X satisfying 0 R A.
For each pair (i,j) consider the operator defined by =
and Ajj(Xm) 0 for m j. Since this operator satisfies 0 $ A,
it follows that Y is invariant under each operator Therefore, for each
n E N the vector subspace Y is invariant under the operator
=
i=1 j=1
It remains to notice that the sequence of operators converges in the
strong operator topology to B. Therefore, 13(Y) Y, and thus the operator
B has a non-trivial closed invariant subspace.5
Corollary 10.69. Let X be a Banach space with a basis. that a
positive matrix A defines a operator on X which is quasi-
nilpotent at a non-zero positive vector. If a operator T: X + X
is defined by a matrix T = satisfying 0 whenever = 0, then
the operator T has a non-trivial closed invariant subspace.
5As a matter of fact, Y is a common invariant subspace for all operators generated by the
weighted matrices satisfying the condition of the theorem. Compare this with Exercise 9
in Section 10.4
440 10. Invariant Subspaces
Exercises
1. Let be a basis in a Banach space X and let
Clearly, the vector space C(U*, X*) equipped with the norm
= sup lf(x*)I1, f E C(U*7X*),
is a Banach space. The Banach space C(U*, X*) was studied first in [212]
and Observe that for each operator T E L(X) the restriction of the
adjoint operator T* to is an element of C(U*, X*).
V. I. Lomonosov [212], inspired by L. de Branges' proof of the
sical StoneWeierstrass theorem [7'3], characterized the extreme points
of the closed unit ball of the norm dual of C(U*, X*). Subsequently,
L. de Branges [75] presented an analysis of these extreme points and
obtained an abstract version of the StoneWeierstrass Theorem. The
Lomonosovde Branges analysis will be employed later on in our charac-
terization of the invariant subspace problem.
As usual, C(U*) denotes the Banach space of all continuous real-valued
(or complex-valued if X is a complex Banach space) functions defined on
Each function a E C(U*) defines a multiplication operator on C(U*, X*) via
the formula
(af)(x*) = a(x*)f(x*), f E C(U*,X*), E
Clearly,
IIaf II
Every pair (x**, x*) E x gives rise to a linear functional
on C(U*,X*) defined for each f C(U*,X*) by
(f,x** x*) = (x** x*)(f) = (f(x*),x**) = x**(f(x*))
Proof. Clearly, a(Y, Y') r8. So, it suffices to establish r8 'r(Y, Y'). To
this end, fix E We must show that the set
{f E Y: = lIf(x*)II 1}
is a 'r(Y, Y')-neighborhood of zero. Consider the operator
R: (X**,cr(X**,X*)) (Y',cr(Y',Y)),
defined by R
0 in X**, then (f, = (f, = (f(x*), 0 for
each f E Y, and so 0. Since is a set,
it follows that the convex circled set D = R(U**) = {x** 0 E }
is a(Y', Y)-compact. So, its polar
Corollary 10.72. The topologies and on C(U*, X*) have the same
closed convex sets.
1)
= f(xfl) = = =0
(2) There exists an operator B E (X) such that its adjoint operator
B* does not belong to the r3-closure in C(U*,X*) of the vector
space generated by the set {aT*: a E C(U*) and T E A).
Proof. Let M denote the vector subspace of C(U*, X*) generated by the
collection of functions {aT*: a E C(U*) and T E A), and let M denote
the closure of M in the topology 'r3.
10. Invariant Subspaces
for each T E A and each a E C(U*). That is, the 'r5-continuous linear
functional vanishes on M. On the other hand, we have
(x** x* = (x** b b*) = (b*, x**)(b x*) = 1,
bind so Lemma 10.73 implies that B*
(2) Pick some B E
(1) such that B* M. Since M is invariant
under multiplication by elements of C(U*), it follows from Lemma 10.73
that there exist E X** and E U* such that
(x** B*) = (B*x*, x**) = 1 and (T*x*, x**) 0
f'or all T E A. Since (B*x*,x**) = 1, it follows that 0 and 0,
and by Theorem 10.12 the algebra A* is non-transitive.
Our next goal is to obtain a similar characterization for the invariant
subspace problem in terms of the norm topology. To accomplish this, we
need to introduce a new class of functions.
Definition 10.75. A function f Y = C(U*, X*) is to be completely
continuous if it for the weak* topology on U* and the norm
topology on
The vector of all completely functions of the Banach
C(U*, X*) will be denoted by 1C(U*, X*).
(2) M
is norm closed.
for all T E A. The former condition implies that K*x* 0, and hence the
latter condition shows that Theorem 10.12 is applicable to A*. Thus, the
algebra A* is non-transitive.
and this shows that K* B* does not belong to the norm closure of Al.
(2) Assume that the operators B and K satisfy the stated proper-
(1)
ties. Again, let Al denote the vector space generated in C(U*, X*) by the
collection of functions a E C(U*) and T E A}. Clearly, Al is
K*B* belongs to 1C(U*, X*) and Al IC(U*, X*).
Since the compact operator K*B* does not belong to the norm closure
of Al, it follows from Theorem 10.76 that K*B* is not in the of
Al in 1C(U*, X*). Consequently, by Lemma 10.73, there exist E X** and
E U* such that
(K*B*, 0 = (K*B*x*, = 1 and (K*T*x*, = 0
for all T E A. Since (K*B*x*, x**) = 1, we see that xO = K**x** 0.
though the given algebra A were not connected with any compactness
erties in any way from the outset!
Exercises
1. If 11 is a compact topological space and X is a Banach space, then
show that the space C(11,X) of all X-valued continuous functions on
equipped with the norm
If 1 = max If (w) II
is a Banach spacewhich is a Banach lattice if X is a Banach lattice.
2. Let X be a Banach space, and let the closed unit ball U* of X* be
equipped with the Establish the following.
(a) If T: X X is a compact operator, then T*: X* (the re-
striction of T* to U*) is a completely continuous function.
(b) The collection of all completely continuous functions X*) is a
norm closed subspace of C(U*, X*), and so it is a Banach space in
its own right.
3. Assume that for a subalgebra A of (X) there exists a non-zero operator
T0 E (X) such that:
(a) T0 commutes with each operator in A.
(b) For some B E (X) its adjoint operator B* is not in the re-closure of
the vector subspace spanned by a E C(U*) and T E A}
in C(U*,X*).
Show that A* is non-transitive. [HINT: By Lemma 10.73, there exist two
vectors E and E U* satisfying
(x,x*) - <6
for each n. Therefore, the identity operator 1* does not belong to the
r8-closure of M and the conclusion follows from the previous exercise.]
5. (Honor [156]) If A is a commutative algebra of bounded operators on
then show that the dual algebra A* of operators on is non-transitive.
= + a), 0
has a non-trivial closed hyperinvariant subspace.
6 Recall that a real number u is said to be a Liouville number if there exists a sequence
of pairwise distinct rational numbers such that (mn, = 1 and [u I< holds
for each ri. For instance, the real number u = 0.110001. is a Liouville number.
Every Liouville number is transcendental.
7A collection of operators on a Banach space is said to be a Volterra collection if each
operator in the collection is compact and quasinilpotent.
452 10. Invariant Subspaces
The major step needed to prove Theorem 10.84 was the verification that
the algebra generated by a multiplicative Volterra semigroup remained like-
wise a multiplicative Volterra semigroup. This verification is very technical
and uses several subtle ideas. After that, the existence of the
variant subspace is a simple consequence of Shulman's Theorem 10.83. For
a complete exposition of Theorem 10.84 we refer the reader to the excellent
monograph by H. Radjavi and P. Rosenthal [273J.
Recall that if X is a finite dimensional Banach space, then the classical
Burnside theorem states the following.
Theorem 10.85 (Burnside). If X is a finite dimensional complex Banach
space, then each proper subalgebra A of (X) is non-transitive.
For a proof of this theorem see [273, Section 1. 2}. It is worth pointing
out that Burnside's theorem is not true for finite dimensional real Banach
spaces. To see this, let X = 2, and take any non-scalar operator T
on X without a non-trivial hyperinvariant subspacewe refer to Exercise 2
in Section 10.1 for the construction of such an operator. Consider the algebra
A = {T}'. It is transitive, since T does not have a non-trivial hyperinvariant
subspace. At the same time a straightforward verification shows that this
algebra A is necessarily proper, since T is a non-scalar operator.
We will mention two elegant ways of extending the Burnside theorem to
infinite dimensional Banach spaces. The first result is due to G. L. Litvinov
and V. I. Lomonosov [209] and it is true for locally convex spaces as well.
Notice that while discussing the transitivity of a subalgebra A of (X) we
can always assume that A is weakly closed, that is, closed in the weak
operator topology (otherwise we can replace A by its weak closure which
remains non-transitive if A were non-transitive).
Theorem 10.86 (LitvinovLomonosov). Let X be a complex Banach space
and let A be a weakly closed subalgebra of (X). If A is transitive and
contains a non-zero compact operator, then A =
7-1 and there exists a normal operator B on N that leaves the subspace H
invariant and A = BIH, that is, the restriction of B to H conicides with the
initial operator A.
Regarding subnormal operators, S. Brown [76] proved the following re-
markable resultthat was viewed as a major achievement following a long
period of partial successes by many mathematicians working on the invariant
subspace problem for Hilbert space operators.
Theorem 10.89 (S. Brown). Every subnormal operator has a non-trivial
closed invariant subspace.
It turns out that various classes of operators on many other Banach spaces
satisfy this equation, which is known today as the Daugavet equation.
In 1965, C. Foias and I. Singer [126] extended Daugavet's result to arbi-
trary atomless and to several classes of operators (including
the weakly compact ones), and in 1966 G. Ya. Lozanovsky [219] established
that compact operators on L1 [0, 1] satisfy the Daugavet equation.
During the next fifteen years dr so these results were left without much
attention approximately until the beginning of the 1980's. At that time a
new wave of interest in this topic surfaced and the Daugavet equation has
been studied by many authors in various contexts. The major results of
these studies will be reported in this chapter. The introduced techniques
vary from purely order-theoretical to analytical and we intend to familiarize
the reader with the majority of these methods. However, this has its price.
Our presentation is not as efficient as it could be, since we prove several
basic results repeatedly to demonstrate different methods. For the most
part these results deal with various classes of operators on and L1 (ii;).-
spaces. These spaces are Banach lattices and we have an option to utilize
their order structures.
We hope that the resulting comprehensive treatment of the currently
available methods on the subject is an acceptable trade-off for the occasional
redundancy of a few proofs. The last section of the chapter demonstrates
how the Daugavet equation, a purely isometric property, can be used to
455
456 -- 11. The Daugavet Equation
fI+TjI
+ -fJ -
=
Taking the limit, we see that 11 + TJJ I + 11Th, and so T satisfies the
Daugavet equation.
Proof. Assume that the vectors u and v satisfy lull + and let
a, ,8 0. By the symmetry of the situation, we can assume that ,8 0.
Then we have
= jcE(u+v)
-
= + Ilvil) - -
=
and the desired equality follows.
The next lemma describes two equivalent norm conditions the first of
which is usually taken for the definition of uniform convexity. As usual, we
denote by (Ix the closed unit ball of a Banach space X.
Lemma 11.6. For a Banach space X the following are equivalent.
(1) For each 0 < 2 there exists some 0 < (5 < 1 such that if
x,yE satisfy IIxyIl , then <1(5.
(2) If two sequences and in satisfy = 2,
then IIxn 0.
Proof. (1) (2) Let two sequences and belong to (Ix and
satisfy IIxn + 2. Now fix 0 < 2 and then pick some
0 <(5<1 that satisfies (1). Next, choosesomen0 suchthat > i(5
for all n ? n0. But then it follows from (1) that < for each
n i.e., = 0.
(2) (1) If (1) were not true, then there would exist some 0
and (Ix with > and > 1 This implies
lIxn + = 2 and hence, by (2), we have = 0,
contrary to > for each n.
Definition 11.7. A Banach space that satisfies either one of the equivalent
statements of Lemma 11.6 is said to be uniformly convex (or uniformly
rotund).
Definition 11.8. A Banach space is called uniformly smooth if for each
0 there is some (5 > 0 such that fJxII 1, 1, and lix < 6
imply tx + llxIt + Mx yll.
458 11. TheDaugavet Equation
Fora proof of the preceding result see Exercises 13 and 14 at the end
of the section. For more information about uniformly convex and uniformly
smooth Banach spaces see [102, 105, 118] and {189, pp. 353369].
For uniformly convex Banach spaces the converse of Lemma 11.3 is true.
Theorem 11.10. If X is uniformly convex and T E (X) satisfies the
Daugavet equation, then 11Th E cT(T).
From Theorem 7.11 we know that the non-zero points of the spectrum
of a strictly singular operator on a Banach space are eigenvalues. Therefore,
the condition that the norm of an operator is an eigenvalue characterizes
the strictly singular operators that satisfy the Daugavet equation.
Corollary 11.12. A strictly singular (in particular, compact) operator T
on a uniformly convex or a uniformly smooth Banach space satisfies the
Daugavet equation if and only if its norm lIT J is an eigenvalue of T.
11.1. The Daugavet Equation and Uniform Convexity 459
Proof. The conclusions of the theorem follow easily from the Spectral
Mapping Theorem. However, we prefer to present an independent and
direct proof. Assume that the Banach space X, the continuous operator
T: X X, and the function f(A) satisfy the stated properties. By
ollary 11.11, we know that there exists a sequence } of unit vectors
satisfying MTxn 0. Using the identity
+
lim 0
flp 00
lim =0.
460 11. The Daugavet Equation
To see this, let 0. Fix some m such that 11Tht < and then
pick some such that II
< holds for all m no.
Therefore, for n no we have
- f(IITI1)x4= -
m 00
i=O i=m+1
<+2
i=m+1
which shows that = 0. Consequently, the
real number 1(11Th) lies in the approximate point spectrum of f(T). Since
111(T) 1 f( 11Th) and the spectrum of f(T) lies inside the closed disk with
center at zero and radius 111(T) we infer that the equality 111(T) II = f( ITH)
must hold. Now Corollary 11.11 guarantees that the operator f(T) satisfies
(1) For each 0 < 2 there exists some 0 < < 1 such that lvll 1
imply <1s.
(2) If C Ux and lim = 1, then lim = 0.
It is well known that there exist locally uniformly convex Banach spaces
that are not uniformly convex. Therefore, it is natural to ask whether the
characterization of operators satisfying the Daugavet equation on uniformly
convex Banach spaces given in Corollary 11.11 remains valid for locally uni-
formly convex spaces. Surprisingly, as the next example shows, the answer
is negative.
11.1. The Daugavet Equation and Uniform Convexity 461
I and + =
For each m define the positive operator Tn: + by
Tn(X,y) = (y,O).
Observe that Tn(Un) = 0, Tn (va) = and IITn IL = 1 for each ri,.
Next, consider the Banach lattice X = (X1 that is, X
is the 2-sum of the Banach lattices and define the positive operator
T(xi,x2,...) = (Tixi,T2x2,...).
The conclusion that the operator T and the space X have the desired
properties will be obtained from the following three statements.
(1) The Banach lattice X is reflexive, fails to be uniformly convex but
is locally uniformly convex.
(2) The operator T satisfies the Daugavet equation.
(3) The norm = 1 does not belong to the spectrum of T.
The reflexivity of X follows from the identity
ilxnynlI2
Therefore, X is not uniformly convex.
The local uniform convexity of X follows from the well-known fact that
the 2-sum of a sequence of locally uniformly convex Banach spaces is locally
462 11. The Daugavet Equation
uniformly convex; see, for instance [218]. Now we shall verify statement (2).
If x = (x1, x2,...) E X satisfies 11x112 1, then note that
00 00 00
=> = <1,
The situation for locally uniformly convex Banach spaces is not as bad
as one might expect from the previous example. For compact operators on
locally uniformly spaces the conclusion of Corollary 11.12 remains true:
Theorem 11.18. A compact operator T: X + X on a locally uniformly
convex Banach space satisfies the Dan gavet equation if and only if its norm
is an eigenvalue of T.
Proof. The "if" part is true for each Banach space and is independent
the compactness of the operator. Indeed, if JjTjj is an eigenvalue of T, then
by Theorem 11.10 we get + TII = 1 + 11Th.
For the converse, assume that T is non-zero and satisfies the Daugavet
equation. By Corollary 11.5, we know that the compact operator S
also satisfies the Daugavet equation, and so sup11x11=1 jx+Sxlj = 1+11811 = 2.
Pick a sequence of unit vectors such that
lim + = 2. (**)
Exercises
1. Show that if the vectors . . , in a normed space satisfy
=
then = holds for each choice of non-negative
scalars a1, . , [Hint: The proof is by induction. For rt = 1, the
claim is trivially true. So, assume the claim true for some n and let
the vectors , . satisfy it Ski = 15k Also, let ii.
= IIxiII +
n+1
= 5k] >(ai ak)xk,
in conjunction with the induction hypothesis, yields
n+1
fl>(ai ak)xkfl
= >(al-ak)IIxkII
lim + + 0
0 a
0
+/3JJvThM)
+ +
= aIIUTh + 0.]
m m
lim
fl+ 00
=0.
k=1 k=1
6. (C.-S. Lin [202]) Generalize Theorem 11.10 by showing that for a uni-
formly convex Banach space X and two operators S, T E (X) the fol-
lowing statements are equivalent.
(a)
(b) There exists a sequence } of unit vectors such that
lim
fl
=1+ + 11Th.
= = II+(I2T)H =2.
This implies TIf < 1.1
(Tf,g) = (f,Tg)
for all f,g E L2(14, where (.,.) denotes the usual inner product in
defined by (f, g) = Establish the following.
(a) Symmetric operators are bounded.
(b) An integral operator T on with a kernel K E x is
symmetric if and only if K(s, t) = K(t, s) holds for x all
(s,t) E X
(c) The integral operator T: L2 [0, 11 L2 [0, 1], defined by
p1
Tx(t) = / sin(st)x(s) ds,
Jo
is symmetric.
(d) Every symmetric operator T: satisfies r(T)
(e) Every symmetric positive operator T: L2 (ii) satisfies the
Daugavet equation, i.e., 1I+T11 1 + IITLI.
[HINT: For (c) note that 112 = (Tf,Tf) = (T2f,f) 112
12. (Milman [242]) Show that uniformly convex Banach spaces are reflexive.
13. Show that a Banach space X is uniformly smooth if and only if for each
E > 0 there exists some 'y > 0 such that for each unit vector u and for
+1(uvfj <2+EIivII.
15. Show that a locally uniformly convex Banach space X always satisfies
the KadetsKIee property. That is, show that in a locally uniformly
convex Banach space a sequence } is norm convergent to some vector
x if and only if x and tixti.
16. Show that each with 1 <p < co is uniformly convex and uni-
formly smooth. Also show that the infinite dimensional L1- and
spaces are neither uniformly convex nor uniformly smooth.
466 11. The Daugavet Equation
(d) For the function v (v) the following statements are equiva-
lent.
(i) is linear.
(ii) = for each v E X.
(iii) The norm of the Banach space X is Gateaux differentiable
at x, i.e., exists in for each v E X.
(iv) There exists a unique supporting functional at x, that is, the
set is a singleton.
(v) (.) is linear and is the only supporting functional at x.
(e) The norm of a Banach space X is said to be Frchet differentiable
at x if the limit Ix+t4- exists uniformly in v in the unit
sphere Sx = {v E X: vii = 1 }. Establish the following.
(cr) If the norm is Frchet differentiable at x, then it is also
Gateaux differentiable at x.
The norm is Frchet differentiable at x if and only if there
exists a (necessarily unique) linear functional E X* such
that: for every > 0 there is some 8 > 0 so that for each
u E X with <8 we have
I
- Mxli _x*(u)1 (**)
18. Show that a Banach space X is uniformly smooth if and only if its norm
is uniformly Frchet differentiable over the unit sphere of X, that is, the
limit
= lim
2 theorem has a long history. For E = C[O, it was proven first by J. Duncan, C. WI. Mc-
Gregor, J. D. Price, and A. J. White [116] and was rediscovered later by J. R. Holub [153]. It
was generalized and proven in its present form by Y. A. Abramovich [4] and K. D. Schmidt [297').
The proof presented here is taken from [41
______________
(Tx)(wo)> ITM -
Clearly, = I and thus x(w0) = 1 or x(w0) = 1. If x(wO) = 1, then
JI+T1I
x(wO) + (Tx)(wo) = 1 + (Tx)(wo)
1+IITJI,
and if x(wO) = 1, then
or III-TIl
Exercises
1. Show that a compact Hausdorif space has an isolated point if and only
if has an atom.
(a) Verify directly (i.e., without using Theorem 11.20) that the operator
T: Lc,40, 1] LOC{O, 1] satisfies the Daugavet equation.
(b) Show that T : L1 [0 , 1] + L1 [0 , 1] satisfies the D augavet equation.
(c) If (g, 1) = 0, then show that T: 1] + 1] does not satisfy
the Daugavet equation for any 1 < p < oo.
5. If ,u is a or-finite non-atomic measure, then show that every finite-rank
operatofbn where 1 p < oc, is disjoint from the identity oper-
ator. (And so, in this case, by Lemma 11.21 every finite-rank operator
on satisfies the Daugavet equation.) [HINT: Assume that
is a non-atomic finite measure space, f E and u e (a), where
1 oc and 1 <q oo satisfy = 1. Toshow IA(fu) =0, it
suffices to establish that IA (fu)(1)
To verify this, assume that some g E satisfies g IA(f'u)(l).
Fix 0. The absolute continuity of the integral guarantees the existence
of some 8> 0 such that IA f(t) d,a(t) holds for each measurable set
A with < Next, choose some n with and then select
pairwise disjoint measurable sets A1,. . , with = for each
i. Now note that g <XAC + [J'A. f(t) + ai for each i. This
implies g Since 0 is arbitrary. the latter shows that g = 0 and
from this it follows that I A (f u) =
6. Show that Lemma 11.21 is false for with 1 <p < 00. [HINT:
Consider, for instance, the rank-one operator T: L2[0,1J + L2[0,1] de-
fined by Tf = where 1 is the constant function one.
Then T is and (since
positive, compact, disjoint from the identity =1
is not an eigenvalue) it fails to satisfyby Corollary 11.13the Daugavet
equation.]
1 and '}
= hAil,
and the validity of (1) has been established.
To prove (2), we begin by choosing a unit vector u E X such that
llu + SuM > IIIx + 811 c. Then we have
f +T1I
= Ilu+ SuM
>
474 - 11. The Daugavet Equation
we have 8 = if T E I((X
P1TR. So, for some operator ideal I,
then it follows from Definition 11.26 that 8 E 1(X). D
Assume first that A1x0 = 0. In this case, we claim that the rank-
one operator 8: X X, defined by Sx = satisfies the desired
properties. To see this, note that
hAil < !IAxoII = IA2xoIl = = ilSxoIl 11811 trAil.
For (2), we have
llxey + 1l(xo,0) +T(xo, 0) II = 1I(xo + I(xo,A2xo)II
= lixolt + IIA2xoII = 1 + tIA2xo!l = I + IIAxoIl
1+ MAIl 1+
IIIx+SIl.
For (3), consider the scheme of bounded operators
R2>X,
where R1x = (x,0), R2(x,y)
= y))xo, and note that 8 = R2TR1.
Now assume A1x0 0. In this case, we define the bounded operator
8: X X by 8x = A1x + where VO
= We claim that
11.3. The Daugavet Property in Banach Spaces 475
S satisfies the desired properties. Let us verify first (1). For an arbitrary
x E X, we have
= lIAixohI + = IIAxohI
R2>X,
where R1x = (x,O) and R2(x,y) = P1(x,y) and note that
S=R2TR1.
Consequently,
weakly compact operators and so, by Theorem 11.30, X also satisfies the
Daugavet property for weakly compact operators. At the same time X is not
lattice isomorphic to an AL- or to an AM-space; see Exercise 3 at the end
of the section. This demonstrates that, indeed, there exist "non-classical"
Banach lattices satisfying the Daugavet property. The first examples of
such Banach lattices were obtained in [5], where finite sums of AL- and
AM-spaces were considered.
Exercises
1. Prove Corollary 11.25.
s ED}.
Show that Eis lattice isomorphic to an AM-space if and only ifM(E) <oo.
3. Let = L1[O, 1] and Fn = 1] for each n. For the Banach lattices
E= and F establish the following.
(a) Neither E nor F is lattice isomorphic to an AL- or an AM-space.
(b) Neither (E F)1 nor (E is lattice isomorphic to an AL- or
an AM-space.
(c) The Banach spaces (E F)1 and (E satisfy the Daugavet
property for weakly compact operators.
The Daugavet Property in 477
From now on, until the end of this section, if is a measure on then we
shall denote the u-essential supremum of a function f by if
Lemma 11.35. If has no isolated points and T: is a
bounded operator, then there exists a non-atomic regular Borel probability
measure on such that =
Proof. For each k choose a function xk E such that iIXk 1 and
Proof. By Lemma 11.35, for each m there exists a non-atomic regular Borel
probability measure on such that = Then the measure
480 11. The Daugavet Equation
Proof. Since Ihn(w)I <oo holds for each w the conclusion fol-
lows from Egorov's theorem and the regularity of the measure
(*)
where:
(c) The series in (*) converges in the strong operator topology, that is,
is norm convergent in C(IZ) for each x E
To show that the series in (*) converges in the strong operator topology, fix
any x E C(IZ), and let Rx = (ai, a2,...) = ==
It follows that
Tx S(Rx) = S(>(Rx, =
sup
wEB n=/c+1
5x0 = J,UTI A
= xEC(11) and xjxO},
= sup +
IxIxo ri=1 m=k+1
inf = 0 (t)
wEG
for each non-empty open subset G of then T satisfies the Daugavet equa-
tion.
Proof. (1) Let }wEcl be the representing kernel of T. Since the range of
T* is separable, there exists an at most countable subset of such that
the collection wE is norm dense in w -.E Therefore, it
is easy to see that the set A = {w E O} coincides with the
set {w E o}. Consequently, A is at most countable.
Exercises 7 and 8 at the end of this section indicate how one can use
Theorem 11.43 to obtain an alternative proof of the fact that each (weakly)
compact operator on an atomless satisfies the Daugavet equation
(Theorem 11.20).
We conclude this section by one more elegant application of Theo-
rem 11.43. Recall that a bounded operator T: X > Y between two Banach
spaces fixes a copy of a Banach space Z if there exists a subspace X0 of
X such that Xo is isomorphic to Z arid T: X0 T(Xo) is an isomorphism.
It should be clear that if is uncountable and a bounded operator
T: * fixes a copy of then the range of T* cannot be
separable.4 Moreover, according to a deep theorem of H. P. Rosenthal [281],
for an uncountable metrizable the converse statement is also true, i.e., if
Exercises
1. Give an example of a continuous operator on C[O, 1] that factors through
CO but that is not weakly compact. [HINT: By a well-known theorem of
lattice isometry.
(b) If also metrizable and has no isolated points, then there exists a
non-atomic regular Borel probability measure on with full support.
[HINT: For (b) mimic the proofs of Lemma 11.35 and Corollary 11.36.1
6. Establish the following separability properties.
(a) If the adjoint of a bounded operator T: X + Y between Banach
spaces has a separable range, then T likewise has a separable range.
(b) If a Banach space X has a separable dual, then X itself is separable.
Also, give an example of a separable Banach space whose norm dual
is not separable.
7. (WeisWerner [330]) If is a metrizable compact space without isolated
points, then show that every weakly compact operator on C(11) satisfies
the Daugavet equation. [HINT: Recall that an operator T E is
weakly compact if and only if
8. (WeisWerner [330]) The BartleDunfordSchwartz theorem [118, p. 306]
states that: A non-empty subset A of = is weakly compact
if and only if it is norm bounded and there exists a measure ,u E
9. (Ansari [37]) Fix q E [0, 1] and consider the rank-one positive operator
T = Sq' on C[O, 1]. That is, Tx = x(q)1 for each x E C[O, 1]. Establish
the following properties of T.
(a) T factors through CO (and hence T satisfies the Daugavet equation).
(b) The operator jT: C[0, 1] k C[0, is not disjoint from ji, where
the operator j: C[0, 1] C[O, is the canonical embedding and
I is the identity operator on C[O, 1].
(c) The adjoint operator T*: C[O, 11* C[0, 1]* is not disjoint from
the identity operator on C[0,
Definition 11.45. Let (X, X') be a dual pair of real vector spaces, and let
C be a weakly bo'anded s'abset of X. Also, let x' E X' and e> 0
be given, and p'at s = (c, x'). Then the slice of C determined by x'
and is the set
= {x E C: (x,x') s
Figure 1
(3) If vo E Y and are unit vectors, then for each 0 <E <
E 1
there exist a unit vector y E Y and some 0 <6 < 1 such that:
(a) Slice(Uy*,y,6) C Slice(Uy*,yo,).
(b) For each ESlice(Uy*,y,6) we have 2.
Proof. We establish only the equivalence (1) (2). The proof of the
equivalence (1) (3) is similar and is left to the reader.
(1) (2) Assume that Yo E Y and E are unit vectors, and let
0<- <-1. Start -by- observing that the rank-one operator- T = -yo in
L(X, Y) satisfies 11Th 1. So, by our hypothesis, + = 1 + 1TII = 2,
and consequently 1IJ* + T*L1 = 2. Fix some linear functional E of
y* y*
norm one satisfying 1
+ > 2 > 0 and (Yo) 0, and let
- x *_
an d
I and 0 <6 < 1.
Fix any x E 6) = {x E x*(x) 1 6}. We claim that
y*(x) + y*(y )x*(x)
2 E. (*)
2__y*(x) Ie>0.
But then y*(yo) 0 yields y*(yo) > 0 and > 0. Therefore,
1 c> 0.
That is, x E ), and so 6) c ).
Now taking into account 0 < 1 and using (*) once more, we get
+ + [1
(1)(2)
for each 0 < < 1. Letting e J, 0 yields JIJ + TJI 2, as desired. D
Proof. We shall show the result only for the slices of So, let E
be a unit vector, let 0 < 1, and consider the slice Also,
fix some 0 < and note that
Pick an arbitrary unit vector u E Slice(Ux, ) and let yo = u. Now
apply the second statement of Lemma 11.46 to yo, and e to get a unit
vector X* and some 6> 0 such that Slice(Ux,x*,6)
and jx + yo JJ 2 for each x e Slice(Ux, x
x u belong to it
follows that the diameter of is greater than or equal to 2
for all 0 < Hence, the diameter of ) is 2, as claimed. D
51n this case, we also say that the linear functional E Z* strongly exposes the point c.
11.5. Slices and the Daugavet Property 491
A strongly exposed
point A strongly exposed
point
Figure 2
From =
it follows that
1, Ii 1. Next pick a linear functional
y*
E that strongly exposes Yo and satisfies y*(yo) = maxYEK y*(y) = 1.
We claim that there exists some 0 < 8 < such that the slice of K
B = Slice(K,y*,8) = {y E K: y*(y) 1 6}
=sup{y*(Tx): xE = max{y*(y): yE K} = 1.
21-f-1>2--2.
Exercises
Proof. If this is not the case, then there exist some 0and a sequence
satisfying < and for each n. Letting
= Bk, we see that 1 1
for each
n, which is impossible.
Note that since we deal with real Banach spaces, any measarable func-
tion x that satisfies xI = is necessarily of the form x = XB\c for
some measurable subset C of B. Clearly, scalar multiples of narrow opera-
tors are again narrow operators. However, the sum of two narrow operators
need not be a narrow operator; see Exercise 7 at the end of this section.
A useful characterization of narrow operators is as follows.
Theorem 11.53. A bounded operator T: E Y is narrow if and only if
for each A E and each > 0 there exists a function x E E such that
and
Jci
Proof. The "only if" part is obvious. We shall prove the "if" part. So,
assume that T is a non-zero narrow operator. We shall show first that for
To see this, fix D e and 77 > 0. We claim that there exist a se-
quence .} of pairwise disjoint measurable functions and a sequence
{ Co, Ci, C2,. .} of pairwise disjoint measurable subsets of D such that for
each n 1 we have:
(a)
(b)
(c)
(d)
(e) IITynII
0< =
=
it follows that = that is, D = Cn. Consequently,
[YI = IYmI = XCn =
The existence of {yi, } and {C0, Ci,. .} will be established by
. .
= \ = = =
Now, for the induction step, suppose that the measurable functions
Yi, Yk and the measurable subsets Ci,.. of D have been chosen
. ,
to satisfy (a), (b), (c), (d), and (e) for each n = 0,1,. ..,k. Since T is a
narrow operator, there exists a measurable subset Ck+1 of Dk = D \
with and some measurable function Yk+i E E satisfying
I = XCk+1 and IITyk+1M < Moreover, we have
=
1 \ 1
Let u = Clearly,
= IITuM
ui and
u(w)=1}andcl2={weA: u(w)=1},then
the last inequality implies /2(d1) We can assume that
/2(d1) /2(d12). Now select any measurable subset d13 of such that
d11
= and consider the function x E E defined by x(w) = 1
if d13 and x(w) = u(w) if w E dl \ d13. It follows that
x d,u = 0, and
<
IITxIl + 11Th' +<
Thus, x is a function with the desired properties.
Corollary 11.54. If T E L1(E, Y) is a narrow operator, then for each A in
and each n there is a measurable subset B of A such that /2(B) = 2Th/2(A)
and the function h = l)xB XA\B e E satisfies
= =
= = XBk,2)'
Proof. This proof is due to Y. A. Abramovich [3, Theorem 3.1]. Fix some
measurable subset A of with > 0, and let be a generalized
sequence of Rademacher functions supported by A. We must show that
holds for all n m and 112 = for each n. Therefore, the sequence
is an orthonormal sequence in L2(,u). Since f E L2(,a), Parseval's
inequality yields <oc. This implies
f = U, > 0,
and the proof is finished. (For a generalization of this result see Exercise 10
at the end of the section.)
are now ready to establish that DunfordPettis and compact operators are
narrow operators.
Theorem 11.57. Each DunfordPettis operatorand hence each compact
operatorfrom E to Y is narrow.
Observe now that the functions x + h and y are disjoint, and hence
ljx+h+yjIi = llx+hlk+ lly[li = 1+ lITxlfi E.
This implies lix + h + Tx[Ii 1 + IITxIli 2. Finally,
II+T1I T)(x +h)jj1 lix + h +TxIli - jIThili
Ix+h+Txlk E 1 +jfTxIIi
Since E> 0 is arbitrary, it follows that 11+ T1I 1 + lIT and so T satisfies
the Daugavet equation. m
Exercises
1. Give an example of a weakly compact operator that is not narrow.
2. Show that each weakly compact operator on L1 (,a) is a narrow operator.
3. If in a scheme of bounded operators E Y _L Z the operator S is
narrow, then show that the operator TS: E + Z is narrow.
4. Give an example of a narrow operator that is not regular.
5. Let T: E Y be a narrow operator. Show that for each A e there
exists a sequence of Rademacher functions } supported by A such that
0.
500 I I The Daugavet Equation
.
The first result of this type is due to V. Kadets [165] and it informs us
that a Banach space with the Daugavet property cannot have an uncondi-
tional basis.
Theorem 11.60 (Kadets). If a Bariach space satisfies the Daugavet prop-
erty with respect to operators, then it does not have an uncondi-
tional basis.
clearly IPAThX QAx1I 0 for each x E X. This implies that llQAIl sup V.
Therefore, sup W sup V. On the hand, it follows from (*) that
sup W 1+sup V. Consequently, sup V = sup W = 00. However, according
to Lemma 1.53, supV must be finite, a contradiction. This contradiction
establishes that X cannot have an unconditional basis.
Since we know that the Banach lattices C[0, 1] and L1 [0, 1] satisfy the
Daugavet property for rank-one operators, Theorem 11.60 provides imme-
diately a new and simple proof of the following well-known fact.
Corollary 11.61. Neither C[0, ii nor L1{0, 1] has an unconditional basis.
such that
(**)
for each k and for all x E Slice(U, 8k).
502 11. The Daugavet Equation
2E
for each x E Slice(U, and for each V E V with = 1. Therefore, by II
(t)
(1- +
(1 + [11(1- Ek)]
n+i n+i
This shows that the sequence is equivalent to the standard unit basis
Proof. Assume by way of contradiction that the closed unit ball of a Banach
space X has a strongly exposed point uO. Pick a linear functional f E
that strongly exposes uosuch that f(uo) = hf hi = 1. Also, fix any non-zero
vector v E X with f(v) = 0.
By our assumption, the rank-one operator T = f v satisfies the Dau-
gavet property, i.e., III+T11 =1 + 1 + lvii. Now take
a sequence of unit vectors such that
Tx =
for all x E Li[0, 1]. We refer to [107] for a comprehensive treatment of the
RadonNikodyrn property that plays an important role in many problems
in analysis.
Corollary 11.64. If a Banach space X has the RadonNikodym property,
then X does not satisfy the Daugavet property for rank-one operators.
Proof. The conclusion follows from Theorem 11.63 by invoking the follow--
ing well-known theorem of R. Phelps: The closed unit ball of a Banach space
with the RadonNikodym property has strongly exposed points. (See
Chapter VII, Theorem 3].)
Corollary 11.65. Every Banach space can be renorrned so that in the new
norm it fails the Daugavet property for rank-one operators.
Proof. It suffices to establish that a given Banach space X can be renormed
in such a way that its closed unit ball has strongly exposed points. To see
that this is possible, take any x0 that does not belong to the original closed
unit ball U and then renorm X by taking for the new closed unit ball the
closure of the circled convex hull of U U {XO}. It remains to be noticed that
is a strongly exposed point of the new closed unit ball.
Exercises
1. Let X be a Banach space and assume that there exists an increasing
continuous function R+ + R and a scalar y> U such that for every
finite-rank projection P we have
505
506 Bibliography
38. A. B. Antonevich and A. V. Lebedev, Spectral properties of operators with shift, Izv.
Akad. Nauk USSR 5cr. Mat. 47 (1983), 915941.
39. C. Apostol, C. Foias, and D. Voiculescu, Some results on non-quasitriangular operators
W, Math. Pures Appi: 18 (1973), 487514.
40. W. Arendt, tlber das spektrum regularer operatoren, Ph.D. Dissertation, University
of Tbingen, 1979.
41. W. Arendt, A. Grabosch, G. Greiner, U. Groh, H. P. Lotz, U. Moustakas, R. Nagel,
F. Neubrander, and U. Schiotterbeck, One-parameter Semigroups of Positive Oper-
ators, Lecture Notes in Mathematics, 1184, SpringerVerlag, Berlin and New York,
1986.
42: N. Aronszajn and K. T. Smith, Invariant subspaces of completely continuous opera-
Ann. of Math. 60 (1954), 345350.
43. W. B. Arveson and J. Feldman, A note on invariant subspaces, Michigan Math. J.
15 (1968), 6164.
44. F. V. Atkinson, The normal solubility of linear equations in normed spaces, Mat. Sb.
(N.S.) 28(1951), 314.
45. A. Atzmon, Power regular operators, Trans. Amer. Math. Soc. 347 (1995), 3 1013109.
46. A. Atzmon, On the existence of hyperinvariant subspaces, J. Operator Theory 11
(1984), 340.
55. H. Bercovici, Notes on invariant subspaces, Bull. Amer. Math. Soc. (N.S.) 23 (1990),
136.
508 Bibliography
90. B. Chevreau, W. S. Li, and C. Fearcy, A new Lomonosov lemma, Integral Equations
Operator Theory 40 (1998), 409417.
91. M.-D. Choi, E. A. Nordgren, H. Radjavi,F. Rosenthal, and Y. Zhong, Triangularizing
semigroups of quasinilpotent operators with non-negative entries, Indiana Univ. Math.
J. 42 (1993), 1525.
92. J. A. Clarkson, Uniformly convex spaces, Trans. Amer. Math. Soc. 40 (1936), 396414.
13L G. Frobenius, Uber Matrizen aus nicht-negativen Elementen, Sitz. Berichte Kgl.
Akad. Wiss. Berlin, 456477, 1912.
132. V. Gantmacher, Uber schwache totaistetige operatoren, Mat. Sb. (N.S.) 7(49) (1940),
301308.
133. I. M. Gelfand, Normierte Ringe, Mat. Sb. (N.S.), 9(51) (1941), 324.
134. 5. A. Gershgorin, Uber die Albrenzung Eigenwerte einer Matrix, Izv. Akad. Nauk
SSSR Ser. Fiz.-Mat. 6 (1931), 749754.
135. N. Ghoussoub, Positive embeddings of L1, (F), and Math. Ann.
262 (1983), 461472.
136. N. Ghoussoub and H. P. Rosenthal, G5-embeddings and quotients of L1, Math. Ann.
264(1983), 321332.
137. W. T. Cowers and B. Maurey, The anconditional sequence problem, J. Amer. Math.
Soc. 6 (1993), 851874.
138. W. T. Gowers and B. Maurey, Banach spaces with small spaces of operators, Math.
Ann. 307 (1997), 541568.
139. J. J. Grobler, Band irreducible operators, Indag. Math. 48 (1986), 405409.
140. A. L. Grornov, Invariant subspaces of weighted permutation operators, Funktsional.
Anal. i Prilozhen 22 (1988), 7576. (Russian)
141. A. Grothendieck, Produits tensoriels toplogiques et espaces nuclaires, iviemn. Amer.
Math. Soc., 16, Providence, RI, 1965.
142. A. Goullet de Rugy, La structure idale des J. Math. Pures et AppI.
51 (1972), 331373.
143. D. W. Hadwin, An operator still not satisfying Lomonosov's hypothesis, Proc. Amer.
Math. Soc. 123(1995), 30393041.
167. V. M. Kadets, Ft. V. Shvidkoy, G. G. Sirotkin, and D. Werner, Banach spaces with
the Daugavet property, Trans. Amer. Math. Soc. 352 (2000), 855873.
168. G. K. Kalisch, On similarity, reducing manifolds, and unitary equivalences of certain
Volterra operators, Ann. of Math. 66 (1957), 481494.
________
Bibliography 513
169. N. I Kalton, An elementary example of a Banach space not isomorphic to its complex
conjugate, Canadian Math. Bull. 38 (1995), 218222.
170. N. 3. Kalton, N. T. Peck, and W. Roberts, Art F-Sampler, London Math. Soc.
Lecture Notes Series, 89, Cambridge Univ. Press, London and New York, 1984.
171. H. Kamowitz, A property of compact operators, Proc. Amer. Math. Soc. 91 (1984),
23 1236.
172. L. V. Kantorovich, On partially ordered linear spaces and their applications in the
theory oflinear operators, Doki. Akad. Nauk SSSR 4(1935), 1316. (Russian)
173. L. V. Kantorovich, Sur les proprits des espaces semi-ordonns Iinaires, C. 1?. A cad.
Sci. Paris Ser. A-B 202 (1936), 813816.
174. L. V. Kantorovich and G. P. Akilov, Functional Analysis, Pergamon Press, Oxford
and New York, 1982.
175. L. V. Kantorovich, B. Z. Vulikh, and A. G. Pinsker, Functional Analysis in Partially
Ordered Spaces, Gostekhizdat, Moscow and Leningrad, 1950.
176. 5. Kaplan, The Bidual of C(X), I, North-Holland, Amsterdam and New York, 1985.
177. 5. Karlin, Positive operators, J. Math. Mech. 8 (1959), 907937.
178. T. Kato, Perturbation theory for nullity deficiency and other quantities of linear
operators, J. Analyse Math. 6 (1958), 273322.
179. Y. Katznelson and L. Tzafriri, On power-bounded operators, J. Funct. Anal.
68 (1986), 313328.
198. P. Lancaster and M. Tismenetsky, The Theory of Matrices with Applications, 2nd
Edition, Academic Press, Orlando and New York, 1985.
199. K. B. Laursen and M. M. Neumann, An Introduction to Local Spectral Theory,
Clarendon Press, Oxford, 2000.
200. N. J. Laustsen, Maximal ideals in the algebra of certain Banach spaces, University
of Leeds Preprint Series, no. 1, 2000.
201. L. Le Cam, Asymptotic Methods in Statistical Decision Theory, SpringerVerlag,
New York and Berlin, 1986.
202. C.-S. Lin, Generalized Daugavet equations and invertible operators on uniformly
convex Banach spaces, J. Math. Anal. AppI. 197 (1996), 518528.
203. J. Lindenstrauss, On operators which attain their norm, Israel J. Math. 5(1963),
139148.
204. J. Lindenstrauss and L. Tzafriri, On the complemented subspaces problem, Israel J.
Math. 9 (1971), 263269.
205. J. Lindenstrauss and L. Tzafriri, Classical Banach Spaces I, SpringerVerlag, Berlin
and New York, 1977.
206. J. Lindenstrauss and L. Tzafriri, Classical Banach Spaces II, SpringerVerlag, Berlin
and New York, 1979.
207. Z. Lipecki, Extension of vector-lattice homomorphisms, Proc. Amer. Math. Soc.
79 (1980), 247248.
208. Z. Lipecki, D. Plachky, and V'J. Thomsen, Extension of positive operators and ex-
treme points I, Colloq. Math. 42 (1979), 279284.
209. G. L. Litvinov and V. I. Lomonosov, Density theorems in locally convex spaces and
applications, Selecta Math. Soviet 8 (1989), 323339.
210. Ju. I. Ljubi and V. I. Macaev, Operators with separable spectrum, Mat. Sb. (N.S.)
56 (1962), 433468.
211. V. I. Lomonosov, Invariant subspaces of the family of operators that commute with a
completely continuous operator, Funktsional. Anal. i Prilozhen 7(1973), No. 3, 5556.
(Russian)
284. P. Rosenthal, Equivalents of the invariant subspace problem, in Paul Halmos: Cele-
brating 50 Years of Mathematics, SpringerVerlag, Berlin and New York, 1991, 179
188.
285. 0. C. Rota, On the representing of averaging operators, Rend. Padova 30 (1960)7
5264.
286. W. Rudin, Continuous functions on compact spaces without perfect subsets, Proc.
Amer. Math. Soc. 8 (1957), 3942.
287. H. H. Schaefer, Spekraleigenschaften positwer linearer Operatoren, Math. Z.
82 (1963), 303313.
288. H. H. Schaefer, Topologische Nilpotenz irreduzibler Operatoren, Math. Z. 117 (1970),
135140.
289. H. H. Schaefer, Topological Vector Spaces, SpringerVerlag, Berlin and New York,
1974.
290. H. H. Schaefer, Banach Lattices and Positive Operators, SpringerVerlag, Berlin arid
New York, 1974.
291. H. H. Schaefer, On the o-spectrum of order bounded operators, Math. Z. 154 (1977),
7984.
292. H. H. Schaefer, On theorems of de Pagter and AndOKrieger, Math. Z. 192 (1986),
155157.
293. E. Scheffold, Das Spektrum von Verbandsoperatoren in Banachverbnden, Math. Z.
123 (1971), 177190.
294. A. R. Schep, Kernel operators, Indag. Math. 41 (1979), 3953.
295. T. Schlumprecht and V. 0. Troitsky. On quasi-affine transforms of Read's operator,
Proc. Amer. Math. Soc., forthcoming.
296. K. D. Schmidt, On the modulus of weakly compact operators and strongly additive
vector measures, Proc. Amer. Math. Soc. 102 (1988), 862866. -
297. K. D. Schmidt, Daugavet's equation and orthomorphisms, Proc. Amer. Math. Soc.
108 (1990), 905911.
298. H.-U. Schwarz, Banach Lattices and Operators, Tebner Texte, 71, Leipzig, 1984.
299. 0. L. Seever, Nonnegative projections on Co(X), Pacific J. Math. 17(1966), 159166.
300. Z. Semadeni, Banach Spaces of Continuous Functions, Polish Scientific Publishers,
Warsaw, 1971.
301. V. S. Shulmari, On invariant subspaces of Volterra operators, Funktsional Anal. i
Prilozhen 18(1984), 84-85.
302. A. An extension of Lomonosov's techniques to non-compact operators,
Trans. Amer. Math. Soc. 348 (1995), 975995.
303. A. Simoni, A construction of Lomonosov functions and applications to the invariant
subspace problem, Pacific J. Math. 175 (1996), 257270.
304. A. M. Sinclair, Automatic Continuity of Linear Operators, Cambridge University
Press, Cambridge, 1976.
305. I. Singer, Bases in Banach Spaces, Vol. 1, SpringerVerlag, Berlin and New York,
1970.
306. 1. Singer, Bases in Banach Spaces, Vol. 2, SpringerVerlag, Berlin and New York,
1981.
307. 0. 0. Sirotkin, Compact-friendly multiplication operators on Banach function spaces,
J. Funct. Analysis, forthcoming.
Bibliography 519
308. V. L. Smulian, Sur la structure de la sphere unitaire dans l'espace de Banach, Mat.
Sb. (N.S.) 9(1941), 545-561.
309. A. Sobczyk, Projection of the space m on its subspace CO Bull Amer. Math. Soc.
,
47 (1941), 938947.
310. A. Spaisbury, Operators not positive with respect to any basis, Quaestiones Math.
23 (2000), 489494.
311. J. ci Stampfii, Compact perturbations, normal eigenvalues and a problem of Salinas,
J. London Math. Soc. (2) 9 (1974/75), 165475.
312. J. Synnatzschke, The almost integral operators in K-spaces, Vesinik Leningrad Univ.
iVlath. Mekh. Astronom. 13 (1971), 8189.
313. J. Synnatzschke, On the adjoint of a regular operator and some of its applications to
the question of complete continuity and weak continuity of regular operators, Vestnik
Leningrad Univ. Math. Mekh. Astronom. 5(1978), 7181.
314. J. Synnatzschke, Uber eine additive Normgleichheit fur Operatoren in Ba-
nachverbnden, Math. Nachr. 117(1984), 175180.
315. A. Szankowski, A Banach lattice without the approximation property, Israel J. Math.
24(1976), 329337.
316. B.-S. Tam, A cone-theoretic approach to the spectral theory of positive linear oper-
ators: the finite dimensional case, Taiwanese J. Math. 5 (2001), 207277.
317. A. Taylor and D. C. Lay, Introduction to Functional Analysis, R. E. Krieger, Malabar,
Florida, 1986.
318. T. Terzioglu, A characterization of compact linear mappings, Arch. Math. (Basel)
22 (1971), 7678.
319. V. 0. Troitsky, On the modulus of C. J. Read's operator, Positivity 3 (1998), 257
264.
320. V. G. Troitsky, Invariant Subspace Problem and Speciral Properties of Bounded Lin-
ear Operators on Banach Spaces, Banach Lattices, and Topological Vector Spaces,
Ph.D. Dissertation, University of Illinois, Urbana-Champaign, Illinois, 1999.
321. V. G. Troitsky, Lomonosov's theorem cannot be extended to chains of four operators,
Proc. Amer. Math. Soc. 128 (2000), 527540.
322. V. G. Troitsky, Measures of non-compactness of operators on Banach lattices,
preprint.
323. S. L. Troyanski, On locally uniformly convex and differentiable norms in certain
non-separable Banach spaces, Studia Math. 37 (1970/71), 173180.
324. Y. V. Turovskii, Volterra semigroups have invarinat subspaces, J. Funct. Anal.
162 (1999), 313322.
325. L. Tzafriri, Remarks on contractive projections in Israel J. Math. 7(1969),
915.
326. L. Tzafriri, An isomorphic characterization of and co-spaces, II, Michigan Math.
J. 18 (1971), 2131.
327. J. Voight, The projection onto the center of operators in a Banach lattice, Math. Z.
199 (1988), 115117.
328. B. Z. Vulikh, Introduction to the Theory of Partially Ordered Spaces, Wolter
Hoordhoff1 Groningen, Netherlands, 1967. (English translation from the Russian.)
329. B. Z. Vulikh and C. Ya. Lozanovsky, Representation of order continuous and regular
functionals iii partially ordered spaces, Math. USSR-Sb. 13 (1971), 323343.
520 -- Bibliography
330. L. Weis and D. Werner, The Daugavet equation for operators not fixing a copy of
C(S), J. Operator Theory 39 (1998), 8998.
331. J. Wermer, The existence of invariant subspaces, Duke Math. J. 19 (1952), 615622.
346. H. J. Zhong, Riesz operators on the spaces have West decompositions, Northeast.
Math. J. 4 (1988), 282288. (In Chinese with an English summary.) MR 90c:47031
ISBN
on rue Web