LINEAR OPERATORS IN SPACES WITH AN INDEFINITE METRIC T. Ya. Azizov
I. S. Iokhvidov Voronezh State University, USSR Translated by
E. R. Dawson University of Dundee
A Wiley-Interscience Publication
JOHN WILEY & SONS Chichester New York
Brisbane
Toronto
Singapore
Originally published under the title Osnovy Teorii Lineynykh Operatorov v Prostranstvakh s Indefinitnoy Metrikoy, by T. Ya. Asizov and I. S. Iokhvidov, Nauka Publishing House, Moscow
Copyright © 1989 by John Wiley & Sons Ltd. All rights reserved.
No part of this book may be reproduced by any means, or transmitted, or translated into a machine language without the written permission of the publisher Library of Congress Cataloging-in-Publication Data: Azizov, T. R. (Tomas IAkovlevich) [Osnovy teorii lineinykh operatorov v prostranstvakh s indefinitnoi metrikoi. English) Linear operators in spaces with an indefinite metric T.Ya. Azizov, IS. Iokhvidov ; translated by E. R. Dawson. p. cm. - (Pure and applied mathematics) Translation of: Osnovy teorii lineinykh operatorov v prostranstvakh s indefinitnoi metrikoi. Bibliography: Includes index. ISBN 0 471 92129 7 1. Linear operators. I. Iokhvidov, I. S. (Iosif Semenovich) II. Title. III. Series: Pure and applied mathematics (John Wiley & Sons) QA329.2.A9913
1989
515'.7246-dc20
British Library Cataloguing in Publication Data Azizov, la Linear operators in spaces with an indefinite metric. 1. Mathematics. Linear operations 1. Title II. lukvidov, I. S. III. Osnovy teorii lineinykh operatorov v prostranstvakh s indefinitnoi metrikoi. English 515.7'246
ISBN 0 471 92129 7
Phototypesetting by MCS Ltd. Salisbury, Great Britain Printed and bound in Great Britain by Courier International, Tiptree, Essex
CONTENTS
Preface 1
The geometry of spaces with an indefinite metric Linear spaces with an Hermitian form Krein spaces (axiomatics) §3 Canonical projectors P± and canonical symmetry J §4 Semi-definite and definite lineals and subspaces in a Krein
1
§2
14
§6 §7
24
of the classes h± Decomposability of lineals and subspaces of a Krein space. The Gram operator of a subspace. W-spaces and G-spaces J-orthogonal complements and projections. Projectional com-
30
pleteness
42 48 64 72
The method of angular operators Pontryagin spaces II". W"'-spaces and G(")- spaces §10 Dual pairs. J-orthonormalized systems and bases Remarks and bibliographical indications on Chapter 1
Fundamental classes of operators in spaces with an indefinite metric
The adjoint operator T`
Dissipative operators Hermitian, symmetric, and self-adjoint operators Plus-operators, J-non-contractive and J-bi-non-contractive operators §5 Isometric, semi-unitary, and unitary operators §6 The Cayley-Neyman transformation §2 §3 §4
Remarks and bibliographical indications on Chapter 2 3
18
space Uniformly definite (regular) lineals and subspaces. Subspaces
§8 §9
§1
1
§1
§5
2
vii
34
78
84 84 91 104
117 133 142 153
Invariant semi-definite subspaces
158
Statement of the problems
158
§1
v
Contents
vi
§2 §3
§4 §5
4
167
subspaces
176
Invariant subspaces of a family of operators Operators of the classes H and K(H) Remarks and bibliographical indications on Chapter 3
185 193
Spectral topics and some applications §1
§2 §3
5
Invariant subspaces of a J-non-contractive operator Fixed points of linear-fractional transformations and invariant
The spectral function Completeness and basicity of a system of root vectors of Jdissipative operators Examples and applications Remarks and bibliographical indications on Chapter 4
Theory of extensions of isometric and symmetric operators in spaces with an indefinite metric
Potapov-Ginzburg linear-fractional transformations and extensions of operators §2 Extensions of standard isometric and symmetric operators §3 Generalized resolvents of symmetric operators Remarks and bibliographical indications on Chapter 5
207
210 210 219 233 243
245
§1
References
Note: The symbols and are used to mark the beginning and end of some section complete in itself. A reference in the text such as 216.3 refers the reader to §6.3 of Chapter 2.
245
252 266 284 286
PREFACE
L. S. Pontryagin's article 'Hermitian operators in spaces with an indefinite metric' appeared in Izvestiya A cad. Nauk, U. S. S. R., more than 40 years ago. The hard war years followed, and probably because of this, the author learnt
of the significance of such operators for the solution of certain mechanical problems only from a small footnote to an article by S. L. Sobolev. We mention that Sobolev's article [I] itself appeared only in 1960. Thus a new branch of functional analysis-the theory of linear operators in spaces with an indefinite metric-takes its origin from 1944, although theoretical physicists had encountered such spaces somewhat earlier (see [XXI] ). We emphasize that we are speaking of infinite-dimensional spaces, since linear transformations in finite-dimensional spaces with an indefinite metric were already being studied (Frobenius) at the end of the previous century, although there has been a revival of interest in them and their applications in our own time.
L. S. Pontryagin's work was continued, above all, by M. G. Krein and I. S. Iokhvidov. They axiomatized Pontryagin's approach to complex spaces with they considered various an indefinite metric, which they called problems about the geometry of such spaces, and they obtained a number of
new facts for operators in H.. M. G. Krein also studied real spaces U, in connection with the so-called Lorentz transformation and also in connection with the theory of screw curves in infinite-dimensional Lobachevskiy spaces (M. G. Krein [3], see also [XV] ). In M. G. Krein's paper [4], he developed an entirely new method, different from Pontryagin's, of proving theorems about invariant subspaces of plus-operators (to use the modern terminology), based on topological theorems about fixed points. I. Iokhvidov [1] suggested the
application of the Cayley-Neyman transformation to the study of the connection between different classes of operators in II,,. All this was subsequently summarized first in I. Iokhvidov's `Kandidat's' dissertation [3], and later in a long article [XIV], written jointly by I. S. Iokhvidov and M. G. Krein, published in 1956. In 1959 its second part [XV] appeared, containing various applications including one to the indefinite problem of moments. vii
viii
Preface
About this time, the theory began to grow not only in depth but also in width. V. P. Potapov became interested in its finite-dimensional analytic aspect, attracting Yu. P. Ginzburg later to working out infinite-dimensional analogues of his results. In these same years it appears that, independently of Soviet mathematicians, R. Nevanlinna in Finland began work on the general
problems of an indefinite metric, and after him E. Pesonen and I.
S.
Louhivaara. Abroad in the fifties and sixties G. Langer (East Germany) joined in the investigations, basing himself on the work of both Soviet and Finnish mathematicians. At about the same time the first survey was published on the geometry of infinite-dimensional spaces with an indefinite metric, carried out by Ginzburg
and I. Iokhvidov [VIII]. This survey already included in part some results from the very beginning of the sixties, which later became years of rapid growth of the whole theory. One after the other appeared papers by Phillips, Langer, Ginzburg, M. Krein, I. Iokhvidov, Naymark, Shmul'yan, Bognar, Kuzhel, and many other mathematicians. The theory found more and more new applications-to dissipative hyperbolic and parabolic systems of differential equations (Phillips), to damped oscillations of infinite-dimensional elastic systems (M. Krein, Langer), to canonical systems of differential equations (M.
Krein, Yakubovich, Derguzov), to the theory of group representations (Naymark), etc. M. Krein's remarkable lectures on indefinite metric [XVII] appeared (unfortunately in a very small edition), and in 1970 was published the book by Daletskiy and M. Krein [VI] in which the methods of indefinite metric found application and further development. At the end of this period the survey by Azizov and I. Iokhvidov [III] was published (1971), and finally Bognar's first book [V] entirely devoted to indefinite metric appeared in 1974. In the sixties to the existing Odessa school (M. G. Krein) and the Moscow
shcool (M. A. Naymark) occupied in this country with the problems of indefinite metric active new centres were added, among which should first be
mentioned Voronezh were the investigations were grouped round I. S. Iokvidov's seminars at the Voronezh State University and the ScientificTechnical Mathematics Institute. Here a large collective of young mathematicians arose (T. Ya. Azizov, V. A. Khatskevich, V. A. Shtraus, E. I. Iokhvidov, E. B. Usvyatsova, Yu. S. Ektov, V. S. Ritsner, S. A. Khoroshavin, and many others), some of whose results are reflected in this monograph. Now, however, when people literally throughout the world are occupied with the problems of indefinite metric, when courses on individual topics have begun to appear ([II], [XX] ), when Azizov and I. Iokhvidov's survey [IV]
carried out in 1979 by order of VINITI (the Institute for Scientific and Technical Information) already included about 400 names, there is a pressing need for an interpretation in the form of a monograph on at least the purely theoretical aspect of the accumulated material. Bognar's excellent book [V] illuminated only part of the theory (as it stood up to 1973). This also applies to the detailed monograph of I. Iokhvidov, Krein and Langer [XVI], containing
Preface
ix
important material, but only on IIx-spaces and operators acting in them. All this provokes an urgent need for a book devoted to the fundamentals of the theory.
The monograph now offered to the reader differs considerably in its contents from the authors' original plan, which was to expound in detail with complete proofs the fundamentals of the theory of linear operators in spaces with an indefinite metric and its applications on approximately the same scale as was
planned in their survey [IV]. However, our plan for such an extended treatment had to be abandoned because of the very limited size of the book, forcing us to almost unavoidable abridgement. And now the question arose: what should be sacrificed? We could in no way sacrifice the rather extensive introductory first chapter setting out the geometry of spaces with an indefinite metric (we remark that in Bognar's work [V] geometry takes up the first half of the book). To do so would have deprived our book of a whole contingent of readers, in particular of students, post-graduate students, and specialists in natural science wishing to investigate the subject-matter but knowing little of its fundamentals. It is also clear from its very title that the book is supposed to shed light sufficiently fully on the central topics of operator theory. We mention at once that after the easy 'warming-up' pace of Chapter 1 on geometry, we allow ourselves a more and more compressed style of exposition in the later chapters, often leaving the reader to think out for himself many of the arguments and their details. With the same purpose many of the auxiliary propositions (sometimes very important ones) have been reduced to the category of exercises and problems, with which each section of the book ends. The range is such that the reader has to go from the quite simple initial problems to increasingly difficult
ones. Some of the problems form important logical links in the text and without their solution it will be impossible to understand some of the proofs. As a result, the aspect which suffered most turned out to be the third part of
our intended plan (cf. [IV])-the application of the geometry and operator theory to actual problems; we touch on this only in Chapter 4,§3, and then only to a very limited extent, and the choice of applications is subordinated entirely to the authors' tastes. As a justification for this may serve the fact that, on the one hand, it would in any case have been quite impossible to satisfy straight away all the wide and
very varied interests of specialists in dissipative hyperbolic and parabolic systems of differential equations, of specialists in the problem of moments, in the problem of damped oscillations of mechanical systems, in the theory of group representations, of geometers, theoretical physicists and others. On the other hand, a number of monographs, extensive articles and surveys dealing
with applications of indefinite metric to the domains mentioned here (and others) have already appeared. It suffices to mention M. Krein's lectures [XVII], Phillips's papers [1], [2], Sobolev's paper [1], Daletskiy and M. Krein's book [VI], the works of M. Krein and Langer [1], [2], I. Iokhvidov
Preface
x
and M. Krein [XIV], [XV], of Naymark [2], of Kopachevskiy [1]-[3], the extensive cycle of articles by M. Krein and Langer [4]-[7] which represent an almost complete monograph, Nagy's book [XXI], etc. But even with such a self-limitation we have been unable to include all the topics of the theory itself (not even in the form of problems). This applies first of all to topics in perturbation theory, various realizations of indefinite spaces, and many details relating to the spaces II,,. In particular, we do not touch on
the theory of characteristic functions for operators in II,,, or questions connected with rigged spaces, variational theory of eigenvalues, etc. The whole theory is set out within the framework of Krein and Pontryagin
spaces, and therefore many generalizations to Banach spaces or simply normed spaces are also omitted. A little is said about them in [IV]. Each chapter is preceded by a short annotation, saving us the need to review here the structure of the book. At the end of each chapter there are remarks and bibliographical indications, but these in no way pretend to be complete. Most of the difficult problems are accompanied by hints, sometimes rather detailed. At the same time, for many of the problems (including the difficult
ones) no hints are given, but instead the source from which they were borrowed is indicated. Such problems have the purpose of extending the circle of readers and of introducing them to the contemporary state of the theory.
The five chapters of the book are divided into sections. All the special notation is introduced in the form of definitions.
In the citation of references Roman numerals indicate a reference to a monograph or survey listed in the first part of the bibliography. The other references, e.g., 'Jonas [2]' refer to the second part of the list-to Russian and foreign journals and other bibliography-arranged in the alphabetical order of the authors' names. The book has been written rather quickly and therefore we have been unable to use the critical remarks and advice of people interested. An exception is V. A. Khatskevich, who read the manuscript of the book and made a number of important remarks for which we are extremely grateful to him. We also thank M. G. Krein, G. Langer, A. V. Kuzhel, Ya. Bognar, N. D.
Kopachevskiy, V. S. Shul'man, and many other mathematicians for their interest which manifested itself, in particular, in the systematic exchange of information. Finally, without the patience, understanding and support shown to us by the members of our families, this book would not have see the light. T YA. AziZOV and I. S. IoKHvmov Voronezh, March 1984.
The unhappy lot has fallen on me of informing the readers of this book that
my dear teacher and co-author, losif Semeonovich lokhvidov, one of the
Preface
xi
founders of the theory of spaces with an indefinite metric, an eminent mathematician and a man of fine spirit, died on 1 July 1984. At this time the
manuscript had already been put into production and it scarcely needed editorial correction later. This fact is due to the deep pedagogic talent and literary mastery of Iosif Semeonovich, who was the first scrupulous editor of the whole manuscript.
T. YA. Azizov
THE GEOMETRY OF SPACES WITH AN INDEFINITE METRIC
1
This chapter consists of ten sections. Its main substance (§§2-8, 10) is devoted to the geometry of Krein spaces-the principal arena of the action of the linear operators studied in this book. The central item is §8, in which the method of
Ginzburg-Phillips angular operators is developed in detail, and some of its applications (for the time being, purely geometrical) are introduced. The whole presentation of the chapter is based on §1, in which we give a short sketch of the theory of linear spaces with an arbitrary indefinite metric (an Hermitian sesquilinear form). The most important particular case of Krein spaces, the Pontryagin spaces Ilk, are studied in particular in §9, though in fact they are encountered in examples much earlier (starting with §4).
In a more compact form than usual the theory is set out of orthogonal projection and the projection completeness of subspaces up to the maximum ones. The question of §10 of the decomposition of a subspace relative to a uniform dual pair also seems to be new. As regards certain generalizations of Krein spaces and Pontryagin spaces, they are illustrated at the enof §§6 and 9 respectively; but preference is given here to those of them (W-spaces, G-spaces) which are used subsequently in the theory of operators in Krein and Pontryagin spaces. §1 1
Linear spaces with an Hermitian form Let Jr .be a vector space over the field C of complex numbers, and let a
sesquilinear Hermitian form Q(x, y) be given on, i.e. the mapping Q:. x . - C is linear in the first argument: Q(X1x,+X2xz,y)_X1Q(xl,y)+X2Q(xz,y)
(xi,xz,yE
iXi,XzEC) (1.1) 1
2
1 The Geometry of Spaces with an Indefinite Metric
and Hermitian symmetric:
Q(Y,X)=
(X,yE.i)
(x,Y
(1.2)
From (1.2) and (1.1) it follows that q(x,µiy1+µ2Y2)=Al Q(X, YO +µ2Q(X,Y2)
(X, Y1,Y2E
; µ1,µ2EC)
-the so-called semi-linearity (or anti-linearity) of the form Q(x, y) in the second argument. Example 1.1: Let . be the vector space over the field C consisting of all finite infinite sequences x= (E1, E2, ..., Sn, ..) of complex numbers (with n = 0
for n > NX) with the natural (co-ordinate-wise) definition of linear operations, and let (a1, az, .., an, ... ] be an arbitrary infinite sequence of real numbers (an E FR, n = 1, 2, ...). We define a form Q(x, y) on j in the following way: if X= (En1n=1E.,Y= (11n1n=iE., then Q(X, Y) = Ei anSn'ljn
(1.3)
n=1
It is clear that Q(x, y) satisfies the conditions (1.1) and (1.2), since for each concrete pair x, y E . the formally infinite sum on the right hand side of (13) reduces to a finite sum having the usual properties of a sesquilinear form. The continual analogue of Example 1.1 is Example 1.2: Let . be the linear space of all finite, complex-valued, continuous functions defined on the whole real axis R. We introduce the form Q by the formula Q(x, y) =
x(t)y(t) du(t)
(x, y E j),
(1.4)
where a(t) is an arbitrary fixed real-valued function defined on fR with bounded variation on each finite interval. A Hermitian form Q(x, y) with the properties (1.1) and (1.2) is called a Q-metric. We find it convenient to introduce a shorter notation for it: [x, Y] = Q(X, Y) 2
(x, y E -J`).
(1.5)
In this section from now on .? is to be understood to be a vector space with
a Q-metric [x, y]. We remark that the form Q(x, y) is, generally speaking, indefinite, i.e. (see (1.5)) the real number [x, x] = Q(x, x) may have either sign. For this reason the Q-metric [x, y] is also called an indefinite metric. We introduce the following classification of vectors and lineals (i.e. linear subsets)
of the space .
;
at the same time let us agree that throughout the rest
of the book the cursive capital letter SL' (possibly with indices: 9+, _V-, Y1, Y°, .1fx(A)) shall always denote a lineal.
§1 Linear spaces with an Hermitian form
3
Definition 1.3: A vector x(E.) is said to be positive, negative, or neutral depending on whether [x, x] > 0, [x, x] < 0, or [x, x] = 0 respectively. It is clear, for example, that for the vector x = 0 (the zero vector) we have [0, B] = 0, i.e., 0 is a neutral vector; but the reverse implication is, in general, untrue: the neutrality of a vector x does not imply that x = 0. The presence or of non-zero neutral vectors depends on the properties of the absence in Q-metric, a point discussed below in paragraph 4. Positive (respectively negative) vectors and neutral vectors are combined under the general term non-negative (respectively non-positive) vectors. Remark 1.4: As follows from the properties of a Q-metric non-negative,
non-positive, and neutral vectors preserve their non-negativeness, nonpositiveness, and neutrality respectively on multiplication by an arbitrary scalar X E C. Positive and negative vectors behave similarly when multiplied by a non-zero scalar X. We denote the sets of all positive, negative, and neutral vectors of a space respectively by .?++(9) _ ++, 10 --(1) = yP --, and °(.?) _°, i.e.
.:YD ++=(xI[x,x]>0)
= (xI [X, X] <0), J1°= (xI [x,x]=0). (1.6)
We also denote by "ffl
++ U
+(J) _ .-p +
°,
P - (.) _
U
°
(1.7)
the sets of all non-negative and non-positive vectors in 9 respectively. It is clear that,sa+ fl -jp -' =Y0; however, it should not be forgotten that any one of these three sets may, depending on the properties of the Q-metric, reduce to (B), and the set .-++ or .? -- may even be void. Definition 1.5: A lineal X (C `) is said to be non-negative, non-positive, or neutral if 2' C .-+, 2' C ?-, or 2' C .4° respectively. All these three types of lineal are combined under the general name of same-definition lineals. It is clear that neutral 2' are simultaneously non-negative and non-positive.
1.6: A lineal 2' is said to be positive (or negative) if 2'C .i++ U (0) (or 2 C .4--- U [0)). Positive and negative lineals are
Definition
included in the common term `definite lineals'. On the other hand, a lineal 2' which contains both positive and negative vectors (2'fl ++ 0, 2' n e - - ;4 0) is said to be indefinite. 3
Suppose two vector spaces .4, and .z are given with the Q,-metric and the
Q2-metric respectively:
[x,Y] = Q,(x,Y) [u,v]2=Q2(u,v)
(x,YE. 1); (u,vE. 2).
1 The Geometry of Spaces with an Indefinite Metric
4
Definition 1.7: The spaces , and 92 are said to be (Q,, Q2)-isometrically isomorphic if there is a linear bijective mapping' T:.4, - :ire such that [x, y], = [ Tx, TY12
(x, y E .5, ).
From this definition it follows, in particular, that the sets ,1' , .ek ,
'00k,
.iPk +, -)Pk - (k = 1, 2) in the spaces ., and .J2 respectively are connected by the ± ± relations fflf+- = T,i' , jO 20 = T.-Yl °, P£ = T.30 i . The same operator T, called a (Q,, Q2)-isometric isomorphism, sets up a one-to-one correspondence
between all the non-negative (non-positive), and, in particular, between the positive (negative), lineals of the spaces .F, and 92 respectively. The concept of a (Q,, Q2)-skew-symmetric isomorphism S between the same two spaces Jr, and 92 is introduced in a similar way. This linear bijective operator 5:,j, i2 acts so that [x, Y] t = - [Sx, SY12
(x, y E .7, ).
We shall frequently have to use these concepts in the following particular case:
Definition 1.8: If .i is a space with the Q-metric [x, y], then the spare J, which is (Q, Q,)-skew-symmetric to the space ., and which coincides with .T
as a set, but which has the Q, (_ - Q)-metric
Q,(x,Y)= -Q(x,Y)= -[x, Y]
(x,YEJ=it)
is called the anti-space of J. 4
1.9
We return to the question of non-zero, neutral vectors.
If 2' (C.) is indefinite, then it contains non-zero, neutral vectors. and [x, x] > 0, but [y, y] < 0. We consider the function ' of
Let x, y E -
the real variable T (- oo < T < oo ) 'P(T) =[(I - T)X + Ty,(I - T)X + Ty],
(i.e., a square trinomial). Since tip is continuous everywhere on IR, and p(0) = [x, x] > 0 but ,p(l) = [y, y] < 0, there is a To (0 < To < 1) such that p(To) = 0, i.e., the vector z = (1 - To)X + Toy (E2') is neutral. Moreover z ;-d 6, because the vectors x and y are linearly independent (see Remark 1.4). Proposition 1.9 cannot be inverted, because non-zero neutral vectors are contained not only in an indefinite lineal but also in any semi-definite (but not definite!) lineal. In this case they play a particular role, which will be explained later in §1.6. 'Throughout the rest of Chapter l we adhere to the classical symbolism for mapping (cf. IXII], for example); i.e., T: v, -.r2 means that the mapping T is defined on the whole of .y,, and the corresponding images lie in .r2. Later, starting in Chapter 2, it will be more convenient to us to treat the symbol T: .y, -.v2 in a rather wider sense (see Chapter 2, §1.1).
§1 Linear spaces with an Hermitian form
5
Definition 1.10: Vectors X, y (E.) are said to be Q-orthogonal if [x, y] = 0. This fact is denoted by the symbol [1] : x [1] y.
5
It
is clear from (1.2) and (1.5) that the relation just now introduced is
symmetric:
x [1] y - y [1] X.
Q-orthogonality of sets .,ll,.4' (C.f) is naturally defined by requiring that x [1] y (for all x E , 11 and y E.4"), and is denoted by ill [1] .N" (a ,N [1] .ill ). In particular, if.'1' reduces to a single vector x (A" _ x)), then we write simply x[1].,11. Definition 1.11:
The Q-orthogonal complement of a set at (C.f) is the set ,11111 = ( x I x[1]-ll).
From the property (1.1) of a Q-metric [x, y] it follows that X11111 is always
a lineal (even if ill is not a lineal). For any sets dl, .4 (C -fl the implication .,10C.4"-.,11[11
(1.8)
is obvious, and so are the relations Jl[11 n._/t/,111 C (.ll +,/v)[11, (.,lln.N)111 J X11111 +.4"111, -11111 [11 3 -11,
(1.9) (1.10) (1.11)
where the sign + between sets means the algebraic sum of the sets..,// and A i.e., the result of addition elementwise of all possible pairs . 11 +.4"
(x+ yI XE 1l,yEAV), and [11[11 = We note that if 21,2'2 (C. are lineals, then (1.9) can be made more precise: 21111 n211 =(2'i+2'2)111.
6
)
(1.12)
Definition 1.12: A vector xo (E2') is said to be an isotropic vector 2' (C.) if
xo* 0and xo[1]Y. It follows from this definition that any non-zero linear combination of isotropic vectors of a lineal .' is again an isotropic vector of 91. Definition 1.13: The linear envelope Lin(xo) of all the isotropic vectors xo (E2') is called the isotropic lineal for 2' and is denoted by 2P°. In other words,
YO =YnY11),
(1.13)
and the equality 2'° _ (0) indicates the absence of isotropic vectors in Y. Speaking rather loosely, we shall sometimes call 2'° the isotropic part of the lineal Y. Definition 1.14: If 2'° = (0), the lineal 9 is said to be non-degenerate, and in the opposite case degenerate.
1 The Geometry of Spaces with an Indefinite Metric
6
It is easy to see that every definite lineal 2' is non-degenerate; but an indefinite ./' can also be non-degenerate. Example 1.15: We return to Example 1.1 and assume additionally that a, _ - 1, and a,, = I (n >, 2). Then the two-dimensional lineal 2' = Lin(el, e2) (C 9), where el = { 1, 0, 0, ... } and e2 = {0, 1, 0, 0, ... J
is indefinite: [e,, e,] = - 1, [e2, e2] = 1 (see (1.3)). At the same time 2' is non-degenerate, for it would follow from the relation xo = Ale, + X2e2 [1] 2' (because of (1.3)) that 0 = Jxo, e,] = X1 and 0 = [xo, e2] = X2, i.e. xo = 0. We leave
the reader to verify that in this example the whole space . is also a nondegenerate (and, as we saw, an indefinite) lineal. Examples (but by no means the only ones) of degenerate ideals are all (;-d {B}) semi-definite lineals which are different from definite lineals. To see this we note that the restriction of a Q-metric [x, y] to any semi-definite lineal 2' is subject to the Cauchy-Bunyakovski inequality: 1.16 If 91 is semi-definite, then I
[x, y] 12 5 [x, x] [Y, Y]
(x, y E 2').
(1.14)
The proof (in the case of .2 being non-negative) follows at once from consideration of the Hermitian form (non-negative for all , n E C) (0,<)
[tx + nY, tx + nY] _ [x, x] I Z 12 + [x, Yl ti + [y, x]kn + [Y, yl I n 12,
for
which the inequality (1.14) is simply the expression of the non-negativity of its discriminant. In the case of a non-positive 2' we arrive at the same conclusion by considering the form (- [x, y] ). From 1.16 immediately follows the proposition: 1.17 In a semi-definite 2' every neutral vector xo (0 0) is isotropic. In particular,
on a neutral 2' the form [x, y] = 0. If (0 * ) xo E 2' and if [xo, xo] = 0, then for any y E 99 we have, by (1.14), 1
[xo, y] 12 s [xo, xol [Y, Yl = 0, i.e. xo [1] Y.
7 Each of the classes of lineals in . described in the preceding sections contains certain special lineals which will be of particular interest to us later.
Definition 1.18: A positive 2' (C.f) is said to be maximal positive lineal if for any positive 2, D 2' we have 21 = Y. Maximal non-negative, maximal negative, maximal non positive, maximal neutral, and maximal non-degenerative lineals are defined similarly.
For all the classes of lineals mentioned one important principle holds; we formulate and prove this principle for positive lineals (the formulation and proofs for the other classes are entirely analogous). Theorem 1.19: (The maximality principle). Every positive lineal Sf' is contained in a certain maximal positive lineal 22max.
§1 Linear spaces with an Hermitian form
7
We consider the set -It = ( 2) of all positive lineals 2' of the space J which contain 91. This set is not empty: 2'E . if. We introduce in it a partial ordering (<) by inclusion, i.e., for 2',, 2'2 E .,if, we put 91, < 22 if and only if 2', C 2'2. We shall show that in elf any chain (i.e. a linearly ordered subset) is bounded above. Indeed, for such a chain J' we obtain the upper bound by forming the union of all elements of the chain, i.e., of all the positive lineals containing J'
which enter into ,'. Because 7 is linearly ordered, this union will again be a positive lineal containing 2', i.e., it will be an element of the set ,/ f, and by construction it contains all the lineals of the chain JZ, i.e., it is the upper bound for 7. By Zorn's lemma, the set A( contains at least one maximal element Ymax, which obviously will indeed be the required maximal positive lineal containing Y. Remark 1.20: It will be shown later that in the particular case most important for us, when the space F with an indefinite metric [x, y] is a so-called Krein space (see §2), realization of the maximality principle for semi-definite lineals does not require the application of Zorn's Lemma, but the construction of the corresponding `maximal object' can be effectively carried out. We also remark that any lineal 9 (C J) can be regarded afresh us a space with Q-metric which contains, therefore, possibly positive, negative lineals and the like, for which in turn the maximality principle operates (locally, i.e. within 2').
8 We consider an arbitrary degenerate lineal 2' (C.i) and its isotropic lineal 2'° (* (0°)). As a well-known, there is an infinite set of decompositions of .' into a direct sum
g_g°+Y,,
(1.15)
where 2', is any (algebraic) complement to 2'°. From the fact that (1.15) is a direct sum, i.e. g° fl 2', = (0), and that all the vectors isotropic for 2' are collected in g° (cf. Definition 1.13), it follows that with any choice of the complement 2', this lineal is degenerate. Moreover, the following hold: 1.21
In every decomposition (1.15) of an arbitrary lineal Y1, where g'0 is its
isotropic lineal, 2', is the maximal (in 4 non-denegerate lineal. 1.22 Every maximal (in 4 semi-definite lineal 2' contains the isotropic lineal g'0 (see (1.15)).
The decomposition (1.15) also leads to the construction of the factor lineal _ 2222°, whose elements (co-sets) will be denoted by z, y and so on. Every class z (E2') is defined, with an arbitrary vector x E z (x E 2'), as usual by the formula z = x + g° (i.e. z = (x + Ay E ,-o). The canonical homomorphism 9' - 2" can be enriched with additional content if by means of it a certain Q-metric is induced from L into 2' according to the rule
Q(z, y) = Q(x, y), where
x E z, y E y.
(1.16)
1 The Geometry of Spaces with an Indefinite Metric
8
It is not difficult to verify that the definition is a proper one, i.e., it does not depend on the chice of the vectors x E z, y E y.
Introducing again the short notation [x, y] = Q(x, y), and [z, y]" = Q(z y for any x1 ER, yl E y we have xi - x = xo E 22°, yl - y = yo E 99°, so that by virtue of the vectors xo and yo being isotopic we have [z, y]^ = [x, y] _
[x + xo, y + yo] = [x1, y1]. Hence, it follows immediately that 1.23 The lineal 21 in the decomposition (1.15) is (Q, Q)-isometrically isomorphic to the factor lineal 2'= 91IYo with the Q-metric (1.16).
9 A decomposition (1.15) of an arbitrary lineal 2' (C.i) into an isotropic lineal M'0 and a non-degenerate lineal _T1 gives rise to the natural question concerning the possibility of the further decomposition of a non-degenerate lineal 21i if it is indefinite, into the direct sum of a positive and a negative lineal. Without going deeply into this difficult problem here (it has given rise to an extensive literature
(see §6 below)), we shall for the moment establish only some comparatively simple facts.
First of all, although by no means all non-degenerate lineals 2' can be decomposed into the direct sum 22 = 2+ + 2- of a positive lineal 2+ and a negative lineal 2- (see Example 1.33 below), the converse is true: 1.24
If 2' (C,f) admits the decomposition
Y_Y++Y-1
(1.17)
into the direct sum of the lineals 2+ (CJP ++U (B)) and _T- (C then 2' is non-degenerate.
U [0)),
If xo E 22 and xo [1] 9?, then decomposing xo in accordance with (1.17) into (x± E 22` ), we obtain
the components xo = xo + xo
0= [xo,xo]= [xo,xo]+ [xo,x+]_[xo,xo]= -[xo,x+ 0= [xo,xo]= [xo,xo]+ [xo,xo]-[xo,xo]= -[xo,xo]. Hence it is clear that [4, xo ] (and also [xo , xo ]) is a real number, i.e. [xo , x6-1 = [xo , xo ], and therefore [xo , xo ] = [xo , xo ], which is possible only if xo = xo = 0, i.e. xo = 8. The Proposition 1.24 can be reversed in a known sense; more precisely, we can prove the following proposition, rather stronger than the converse of 1.24. 1.25
Let
the
non-degenerate
lineal
2
be
the
algebraic
sum
2 = Y+ + 2- = (x+ + x I x± E Y'± 1, where ..y'+ C ,0 ±. Then 2 = 2+ + land M'± are maximal in .1'± (2') respectively. In particular, if either of 22± is definite, then it is a maximal definite lineal in Y. Supposing, for example, that there are 9+ C .e+(21), 2+ D Y+
§1 Linear spaces with an Hermitian form
9
we obtain i.e. x+ E 2+\2'+, 9+ = x+ + x(x± E 2±), and x- = z+ - x+ E 2+ C .*°+, hence [z, z] = 0. But then xo = x = 2+ - x+ is an isotropic vector both in 2- and in (see (1.17)); in particular xo [l] 2±, i.e. xo = z+ - x+ [1] 2, hence g+ = x+ E 2+, which contradicts the choice of
+
g+
0
In view of Proposition 1.25 the result 1.24 can in turn be regarded as a simple consequence of a more general fact:
1.26 Every 2 (C.)
which
contains
a
lineal
2+ C Y' ++ U (0)
(2'- C .? - - U (0)) and which is maximal in 1 + (2) (in I - (2)) is nondegenerate.
If in 2 there is an isotropic vector xo, then, as a consequence of its neutrality, we have xo 0 Y+ (respectively xo E 2- ), and Lin (?+, xo) (resp. Lin{2-, xo)) is a non-negative (non-positive) lineal which is contained in
and which contains Y+ (27) as a proper subset. We have obtained a contradiction. A decomposition (1.17) in which 2+ C
±
, as soon as it exists, generates
two linear projectors P± which relate to any vector x E 2 its components x± = P±x (E.?`-), where x+ + x- = x. In other words, P+ + P- = Ir is the identity operator in 2', and P+ P- = P- P+ is an operator annihilating Y. If 2 (C 2` ), admits a decomposition (1.17) in which 2- (Y+ ) is a definite lineal, and if Y- C -°+ (2) (2- C <-- (2)), then the mapping P+: ± Y' 2+ (P-: 2' - 2-) is injective. For arbitrary 2+ C 0 these mappings Lemma 1.27:
are also injective for 2' C 0++(Y)U(B) (2' C - (2')U {B)). Suppose, for example, that 2" C ' (2) and that for some x E 2' we have P+ x = B. Then X= P+ x + P- x = P- x E 2-; hence x=0. In the case 2' C 3++(2)U (0), if for XE 2' we have P+x= 0, then x = P+ x + P- x = P- x = 0, because P- x is non-positive, and x is positive or equal to zero. Corollary 1.28: Under the conditions of Lemma 1.27 for maximal nonnegativity (maximal non-positivity) in 2 of the lineal it' C jP+ (2')
(2' C I- (2)) it is sufficient that P+2' = i'+ (P-.' = Y-). Suppose, for example, that 2' C 41 +(2') and that P+2' = M+. If also 2, C JP + (91) and 2, D Y', then 2, = Y'; for otherwise for any x, E 2,\M" we would have P+ X, = xi E 2+ and at the same time x, = P+x for some x E 2' (C 2', ), i.e., P+ (x, - x) = 0. But x, - x E 2, and therefore by Lemma 1.27 x, = x E Y', contrary to the choice of x,. 10 We shall return later in paragraph 11 to decompositions of the form (17) considering them in another light, motivated by the concepts introduced in the present paragraph.
1 The Geometry of Spaces with an Indefinite Metric
10
Definition 1.29: 9'1,12 (C.f) are said to be skewly linked if 99, n y2 -1 = 9''2 n 99,11' = (0). To indicate that t1 and 22 are skewly linked we shall write 9', # 912. 1.30 In order that Y1 and 9'2, of which at least one is neutral, shall be skewly linked it is necessary that two conditions hold: a) 2, U Y2 = (0), b) Y1 + Y2 be non-degenerate. If 91, and 912 are both neutral, these conditions are also
sufficient.
Suppose, for example, that 9'1 C jP°. If 9'1 # Y2, then since (see Proposition 1.17) Y1 C 9 jl], condition a) follows from 9'2 n 21lJ = (0). Further, if If furthermore xo E T, + 92, then xo = x1 + X2 (xk E 91k, k = 1, 2).
so that xo [1] 21 + 92, then, in particular, xo E ?H and x, E £, C x2 E Y11] also. But X2 E 912, and so x2 = 0. Hence x, = xo [-L] E2, i.e., x, E Y, n Y F, and therefore x1 = 0. Thus xo = 0, and condition b) has been proved. Conversely, suppose that 9'1, 992 C .?° and that conditions a) and b) hold. We note that in this case condition a) automatically follows from condition b),
which can be replaced by the even weaker condition b') the algebraic sum 991 + Y2 = [x, + x2 I x, E 2,, x2 E 12) be non-degenerate. Since now 2?, C 9'i
and Y2 C -LL, any non-null vector from Y, n Y4' or 9' fl 991111 would be isotropic for 91, + ?2i contrary to condition b'). The simplest example of skewly linked lineals are two arbitrary, onedimensional lineals 9', = Lin(x1) and 22 = Lin(x2), if [x1, x2] 0 (x,, x2 E . ). The fact that in this example the dimensions of Y1 and Y2 are the same is not accidental, as the following lemma shows.
If 9', # 212 and 0 < dim Y1 = m < oo, then dim .9', = dim 22 and for any basis (e;)(" in Y1 a basis (fk)i" can be chosen in 2'2 such that [e1, Al = Sjk (k, k = 1, . . ., m). ('Q-biorthogonality'). Lemma 1.31:
Choose an arbitrary basis (e;)1' in 51 and suppose at first that dim 9'2 > m. Let x1, . . ., x,,,, xn,+1 be linearly independent vectors in 992. Consider the system of m linear homogeneous equations ,[xi,ee]
(j=1,2,...,m) (1.18)
in t°1,
the m + 1 unknowns
E,n+1;
it rhas a non-trivial solution
contrary to , 1, the condition. Therefore dim 2'2 < dim Y, = m, and by the same argument dim Y, < dim 9'2i i.e., dim 9', = dim 9'2 = m. Further, since 9'1 E 9'2111 = (0), a vector fl in Y2 can be found such that [el, fl] = 1, and if m = 1, then the construction of the Q-biorthogonal bases fell and (f,) is complete. But if m > 1, then we first choose (0 ;6) f, E 9'2 = [e,,,, f,] = 0. This is possible, since such that [e2, f,] = [e3, f l = m = dim q2 > dim Lin (e2, e3, ..., e,,,) = m - 1 and the construction of f, is S
ro,n, ron+ ,
i.e, the vector (0 ) xo = Zk'= 1 Skxk E Y2 n
§1 Linear spaces with an Hermitian form
11
carried out by the method used at the beginning of the proof. Since y'2 n ,M;1] = (B), so [e1, fl] # 0 and we may assume [e,, f,] = 1. Then we choose a vector (0 #) f2 E SP2 such that [e,, f2] = [e3, f2] = _ [e,,,, f2] = 0 and [e2, f2] = 1, and so on until the system (fk)i' is completely constructed. Linear independence (i.e., basicity in 2'2) of this system easily follows from the condition of Q-biorthogonality. Corollary 1.32: = :t . o1 +
If £, # 22 and 4, 02 are finite-dimensional,
then
-T2[1]
When Y1 = Y2 = 10), the assertion follows from the fact that X111 =j.
But if (see Lemma 1.31) dim 2', = dim 2'2 = m > 1, then, having chosen Q-biorthogonal bases (ee)l" and (f;)1" in M1 and 2'2 respectively, for any xE.
[x, f;] e;. Then xt E 21 and x - x, [1] fk (k = 1, 2, ..., m), i.e., x2 = x - xi E .211. we put x, = F,; n=
11 The proof in Lemma 1.31 of the coincidence of the (linear) dimensions of skewly linked 2, and M2 is based essentially on the finite dimensionality of 99,
and M2. As regards infinite-dimensional', and ?2, their linear isomorphism does not always follow from -'1 # 2'2. Example 1.33: Let . be a linear space of infinite-dimensional (in both directions) finite (on the left) sequences x = _m, , = 0 for j < jo(x). We introduce an indefinite metric in by the formula [x, yl =
l= -ao
(1.19)
kA-i-1,
where
x = (;); -oo,
Y= (?ii)i - E .
.
We now consider .1 (C.F) consisting of vectors x with t-, = _2 = . . . = 0. It follows from (1.19) that for such vectors we have [x, x] = 0, i.e. °, C .M°. Similarly we define 22 as the set of all y (E-fl with rlo ='q1 =
= 0. Again from
(1.19) we have [y, y] = 0, so that 22 is also neutral. It is clear that: a) 99, n M2 = 10), and b) Y, + 22 = .9 is non-degenerate, and so by 1.30 M1, # t2 At the same time 9?1 and 9?2 are not isomorphic, because 2, is isomorphic to the space of all sequences (o. 1.... ), but 2'2 is isomorphic to the space of all finite
sequences (77-1i77-2....), which has, as is well known, a smaller (namely, a countable) linear dimensionality.
At the same time we show that . is an example of a non-degenerate space which does not admit a decomposition of the form (1.17). For, suppose that . =.+ + M-, where Y+ C e°++U (B), and M- C .j°-- U 10). (EM'+) Consider the mapping T2: 91+ - M'2, which to every vector x relates the vector T 2 x = ( . . . , t_2, _ 1i 0, 0, ...) (E22). This linear mapping is
1 The Geometry of Spaces with an Indefinite Metric
12
injective, because by (1.19) the equality T2x = 0 implies the equality [x, x] = 0, and so x = 0 (because 2+ is positive). The injectivity of the mapping T2: 2- Y2 is proved similarly. But this means that the linear dimensions of the lineals 2+ and Y- do not exceed the linear dimension of the lineal 2'2, and therefore they are not more than countable, which is impossible because their direct sum 2+ + k = = 91, + 22 is isomorphic to the lineal 91, of higher dimension. We point out that the foregoing argument contains essentially the proof of the following (abstract) theorem: Theorem 1.34: If a non-degenerate spare .f with an indefinite metric is the direct sum = 2, + 22 of two non-isomorphic lineals 2',, 2'2 C .?o, then it does not admit decomposition into the direct sum J = 2+ + 2'- of two definite lineals
2+ and -T-.
The proof is the same as in Example 1.33 if we consider that the decomposition . = 2', + 212 generates projectors Pk: . 2'k (k = 1, 2), P, + P2 = Ii, and use again that one of these two projectors which correspond to the ideal (let us say, 2'2) of lesser dimension (than 2', ).
12 In conclusion we return to the situation in paragraph 3 when there are two complex linear spaces f i and W 2 with a Q,-metric and a Q2-metric [ , ], and -,4'2. However, P, , ]2 given on them respectively, and a linear mapping T:. instead of the stringent requirements of isomorphism and (Q,, Q2)-isometricity here a more `liberal' condition will be imposed on T which, in the notation of
paragraph 3, reads T?o, C -)02 +.
Lemma 1.35: Suppose that the Ql-metric in the space Jr, .is known to be indefinite, the Q2-metric in .4'2 is arbitrary, and the linear mapping T. J, -4,2 has the property T. 1 ° C -,02'. Then, for all y E P i and z E .01 [ TY, TY] 2
< [ Tz, Tzl2
(1.20)
[z, zl ,
[Y, Y] I
Restricting ourselves to `normalized' vectors y: [y, y],
1
and
[z, z], = 1, we rewrite the relation (1.20) as - [Ty, Ty]2 < [Tz, Tz]2, and we assume the contrary. Namely, suppose that for some yo E .s , z:
with
[yo, yo] I = - 1
and
Zo E .,
with
[Zo, zo]i = 1
we
have
- [Tyo, Ty]2 > [ Tzo, TZo12. We consider the vector xo = eyo + zo where c (I e I = 1) is, for the time being, an arbitrary parameter. We have [xo, xo], = 2 Re(e[yo, zo],),
[Txo, Txo]2 < 2 Re(e[Tyo, Tzo]2)
(0 < p < 2ir), so, by choosing the argument p suitably, we But since e = can always arange that Re(c [yo, zo] 1) = 0 and Re(c [ Tyo, Tzo]2) < 0. For such
an a we obtain [xo, xo], = 0 and [Txo, Txo]2 < 0, which contradicts the condition T.901 C -i'i'-
§ 1 Linear spaces with an Hermitian form Corollary 1.36:
µ+(T) = inf
13
Under the conditions of Lemma 1.35 the finite limits [Tx, Tx]z
x E .l t
1
[x, x] 1
(> -ao ), it _(T) = sup xE
[Tx, Tx]z [x, x] 1
exist. Moreover, µ_(t) < µ+(t) and for any it with µ_(T) < µ < µ+(T) the inequality
[Tx,Tx]z>µ[x,x]1
(1.21)
holds for all x E .i.
For the finiteness of the limits µ+ (T) and µ_ (T) and the inequality µ_ (T) <,A+ (T) follow immediately from (1.20). Further, for x E : ' ? the inequality (1.21) follows from the condition T?° C P2'. In the remaining cases when x E sP 1 + U .0 i -, it is easily seen from the chain of inequalities (cf. (1.20))
[Tx,Tx]2<14<
sup (µ-(T)=) XE.'
[x, x] I
[Tx,Tx]2
inf
XE.'i -
[x,x]I
(=µ+(T))
0 Exercises and problems 1
Verify that for the form [x, y] defined in para. I the normalization formula 1
1
i
[x, Y]=4[x+y,x+y]-4[x-y,x-y]+4 [x+iy,x+iy]
-- [x-iy,x-iY]
(x,YE.f)
holds. 2
Consider the linear space f = C[-1, 1] of all complex-valued continuous functions defined on the interval [-1, 1]. Verify that the form [x, y] = f '
x(t)y(- t) dt
(x, yE.f)
(1.22)
is Hermitian and defines an indefinite metric on .f . Give examples of positive and negative vectors in J r. 3
If a space .f with Q-metric [x, y] contains at least one positive (negative) vector, then j w'= f++ + f++ (f = .)P- - + f - - ), where the symbol denotes the algebraic sum of sets (see M. Krein and Shmul'yan [21). Hint: For a positive vector x+ E .f (for instance) and an arbitrary x E .f consider the quadratic form [x + ax+, x + ax+] for large a E R.
4
In order that f (see Exercise 3) should be indefinite it is necessary that not one of the sets .?++ ,--, -P- should be a lineal, and it is sufficient that at least neither of the sets .f°+,.jP - should not be a lineal. (Cf. M. Krein & Shmul'yan [2].) Hint: Use the result of Exercise 3. f+,
1 The Geometry of Spaces with an Indefinite Metric
14 5
Prove that for any set if (C.fz) it is always true that .,I1111 Il] l1' = .,!/l1) (E. Scheibe [1]).
6
Prove that the relation (1.12) holds not only for lineals but for any sets It,, . //Z (C sr)
provided only that they both contain the vector 0. 7
If 2' C .>a°, then 2(11 [ll C jP° ([V]).
8
If 2' is degenerate, then M'Ill [l is also degenerate ([V] ). A similar proposition for non-degenerate lineals is false.
9
If 2' C+, then Y[1] [l may turn out to be indefinite (and may even coincide with the whole space) ([V] ).
10
Prove the converse of Proposition 1.17. Hint: Use the decomposition (1.15) and Proposition 1.9.
11
Obtain the converse of Proposition 1.21.
12
Verify that the inequality (1.21) may fail when µ 0 [µ_ (T), µ+(T)]; construct an example.
§2 1
Krein spaces (axiomatics) We consider a linear space . with a Q-metric Q(x, y) = [x, y]. Suppose
that i admits decomposition into the direct sum of a positive (.+) and a negative ( ) lineal:. _ + -- . -, from which it follows by 1.24 that the whole space Jr .is non-degenerate. We take the following step and suppose that .J+ [1] . -, and so we can write "g =
+ [+]
-'.
(2.1)
where the symbol [+] denotes the Q-orthogonal direct sum. Definition 2.1: The decomposition (1.21) is caled a canonical decomposition of the space .4.
The lineals .;+ and j- in (2.1) are pre-Hilbert spaces: f + with the scalar product (a positive-definite form) [x, y] (x, y E ii + ), and .4- with the scalar product (- [x, y]) (x, y E . -' ). Definition 2.2: A space t with a Q-metric [x, y], which admits a canonical decomposition (2.1) in which + and Jr- are complete, i.e. Hilbert, spaces relative to the norms 11 x II = [x, X]'12 (x E 9 +) and 11 X I I = (- [x, x] "Z) (x E .
)
respectively, is called a Krein space.
Remark 2.3: Here that case in which one or both of the spaces -J r± is finite-dimensional is not excluded. §9 is devoted entirely to this important particular case. But if, in particular, .4- = [6) (or .:F+ = 10)), i.e., the whole of .t is definite, then this Krein space is simply a Hilbert space with the scalar
product [x, y] (or - [x, ] ).
§2 Krein spaces (axiomatics)
15
In considering a Krein space, we shall from the very first time start from the
2
fixed canonical decomposition (2.1) which features in its definition (2.2), although in the case where . is indefinite, as may easily be understood (see theorem 8.17 below), this decomposition is not unique. A canonical decomposition (2.1) enables a (positive definite) scalar product to be introduced into the whole Krein space (i.e. for all x, y E .) according to the formula
(x, y)= [x+,Y+]- 1X-' Y- 1,
x=x++x-,
Y=y++y ;
x+,Y+EI+;
x-,yE ."-.
(2.2)
In particular, for vectors u, v E .J+ we have, clearly (u, v) = [u, v], and for u, v E .: - we have (u, v) = - [u, v]. But if u E J+ and V E . -, then it follows from (2.2) that (u, v) = [u, 0] - [0, v] = 0, i.e.. + 1 -, where the symbol 1 denotes, as usual, orthogonality relative to the scalar product (2.2). Thus, the space with the scalar product (2.2) can be regarded as the orthogonal sum of the two Hilbert spaces J+ and . -, from which follows 2.4
The space .t is complete, i.e. it is a Hilbert space relative to the norm
IIxII =(x,x)tn
(2.3)
generated by the scalar product (2.2). Remark 2.5: Since our main purpose later will be to study linear operators acting in Krein spaces, the topology of this space (and it is defined by the norm (2.3), i.e., by the scalar product (2.2)) is important for questions connected with the boundedness and closure of operators, with their spectral theory, and
so on. At the same time the Definitions (2.2) and (2.3) may create the impression that the topology about which we are speaking depends essentially on the actual choice of the canonical decomposition (2.1). Later (see Theorem 7.19) it will be proved that this impression is erroneous: all norms defined by different canonical decompositions of the form (2.1) turn out to be equivalent.
3
In connection with the above definitions it is expedient to distinguish Krein
spaces from all spaces .
with an indefinite metric by means of a special
notation. For this we shall adopt the traditional notation for Hilbert spaces (as
we saw in paragraph 2, a Krein space is a Hilbert space), namely .. The canonical decomposition (2.1) will now be written in the form
.W_.0+[+]
(2.4)
we recall, moreover, that not only is + [I] . -, but also .,Y+ 1 .W-, i.e. the decomposition (2.4) could also be written in the form =
,+ O+
.
1
where O+ is the symbol for a (Hilbert) orthogonal sum.
(2.5)
1 The Geometry of Spaces with an Indefinite Metric
16
Returning to formula (2.2) we can, in a known sense, invert it, i.e., express the metric (the Q-form) [x, y] in terms of the scalar product (2.2). Keeping for arbitrary vectors x, y E . the notation x±, y± for their components lying in .W+ respectively (in accordance with (2.4) or (2.5)), we have [x, Y] =
[x+
+ X-' Y+ + Y ] = [x+, Y+] + [x , Y ] = (x+, Y+) - (x , Y ),
or finally
[x, Y] = (x+, y+) - (x-, Y-).
(2.6)
[x,x]=llx+II2-IIx-II2
(2.7)
In particular,
From the formula (2.7) we obtain 2.6 Positivity, negativity, or neutrality of a vector x E ,f are equivalent to the relations 11 x+ 11 > 11 x
11,
11 x+ 11 < 11 x
11, or 11 x+ 11 = 11 x
11
respectively.
The results of paragraph 3 show that a Krein space can be looked on as an arbitrary Hilbert space .' decomposed into the usual orthogonal sum (2.5) of the two subspaces W+ and .* - (, + 1 W-) and equipped in addition to the original Hilbert metric (i.e., the scalar product (x, y)) with the additional indefinite metric [x, y] given by the relation (2.6). We suggest to the reader 4
that, as a simple exercise, he should convince himself, that with such a definition of the form Q(x, y) = [x, y] the subspaces Ye+ and .-W- turn out to be positive and negative respectively relative to it, that they are Q-orthogonal (,W+ [I] ,e-), and therefore the decomposition (2.5) is simultaneously also a
canonical decomposition of the form (2.4). Moreover, the original scalar product, as can easily be calculated, is again expressed in terms of the form [x, y] by means of the relation (2.2). In the next section it will be shown (see §3, para. 5) that the point of view presented here on Krein spaces can be further modified and that to this concept can be given an extremely flexible form most convenient for the theory of operators.
Exercises and problems 1
2
The anti-space (see Definition 1.8) of a Krein space is again a Krein space. Suppose that in a canonical decomposition (2.1) of a space . at least one of the terms .t + or .4- is not complete relative to the norm 11 x 11 _ I [x, x] 11,12; we complete s± * relative to this norm to Hilbert spaces .t . Show that the orthogonal sum J ®j is a Krein space which contains i as a dense part. We obtain the same result if we at once complete the whole of .s relative to the norm (2.3).
§2 Krein spaces (axiomatics) 3
17
In Exercise 2 of § 1 the space .J admits a canonical decomposition J = JT+ [+]. -, where J+ and J- are the spaces of all continuous even and odd functions on [ - 1, 1] respectively. Completing it in accordance with the procedure considered in the L2[_ 1, 1] with the indefinite metric (1.22), previous Exercise leads to the space . = in which the Riemann integral should be replaced by a Lebesgue integral. The interval
(- 1,1) can, of course, be replaced by any symmetric interval (- a, a). 4 We modify Example 1.1 by considering the space .F of infinite (in both directions) and finite (also in both directions) sequences x = (..., E-2, t-1, S1, 2, ... ), with co-
ordinatewise definition of the linear operations, and introducing for x, y E o the indefinite metric
(V= (?Ik)k=-W)k
Q(x, y) = (x, Y] = Z k k - L: k=1
(2.8)
E-k71-k.
k=1
Find a canonical decomposition of the space . and show that completion of .4, in the manner of Exercise 2 leads to a Krein space which is Q-isometrically isomorphous to 12 Q+ 12, with the indefinite form (2.8) ([XV]). 5
In Example 1.2 we restrict ourselves to a finite interval [a, b] C R and we denote by w(t) the total variation of a function a(t) on the interval [a, t] (a <, t < b). Then the indefinite form 6
x(t)y(t) da(t)
[x, Y] _ a
(x,YE Lb))
b) space of all w-measurable and w-summable squared functions the structure of a Krein space. Find its canonical decomposition (I. Iokhvidov and Ektov [2] ). Hint: Use Hahn's decomposition theorem (see, e.g. [VII], [VIII], §5, Theorem 1). gives in the
6
Generalize the example in the preceding exercise to arbitrary L2(S,
µ)-spaces (see
[XII] ), where µ = µ+ - µ- is a measure with bounded variation h u I = µ+ + µ_ ( [XII] ). 7
A model of a real Krein space (see, at the end of Chapter 1, the remarks on §1, paragraphs 1, 2) is the orthogonal sum . Y = ,Y+ (D .W- of two real Hilbert Y+ and
.- with a form
Q(x, Y) = [x, Y] = (x+, Y+ )+ - (x-, y-)-,
are the scalar products in k'
where x = x+ + x-, y = y+ + y-, and and W- respectively.
77
x+
0 Fig.
I
1 The Geometry of Spaces with an Indefinite Metric
18
Show that in the particular case when .W' = W- =12 (a real l2), such a Krein space
is Q-isometrically isomorphic to the completion relative to the l2 norm of the analogue of the space.:O'in Exercise 4 for real sequences with a real form (2.8). Verify
that this analogue for the case dim .0' = dim Y- = 1 is Q-isometrically isomorphic to the Cartesian plane IR2 (see Fig. 1) with rectangular Cartesian axes (.xr', W- ), with vectors x , = (E1, n1), X2=(102,772), and with the forms: indefinite form Q(x1, X2) = [XI, X2] =tEI¢2 - nI'h2, and the Hilbert form (x1, x2) = 5142 + 701712 (the scalar product).
In the case of real ,Y', M- with dim Ye' = 1, dim -W- = 3, the Krein space fP = .Y' Q+ .Y- is Q-isometrically isomorphic to the Minkowski space of the special theory of relativity with the Q-metric (see (2.8)) [3
/ Q(x,Y)= [x, Y] = E0170 - L, Sk7lk tt
k=1
8
Let -*'=,,Y' ( 1- be a Krein space and let x± E.,Y±. Then the conditions a) x± 1 y and b) x± [l] y are equivalent for any y E Jf. Hint: Use formula (2.2).
9
Under the conditions of the previous exercise
(XO[L]x,xolx) a IXo J- X',XO i X ) for any
xo=xo +xo
and x=X'+x- in.
(XO,x±E..Y±).
Hint: Use formula (2.2). 10
Let 97 be a subspace of a Krein space, and let 97° be its isotropic part. The space
2' = Yd 97° will be Krein space if and only if 91 admits the decomposition 9' = Se [+] 99' into the direct sum of the isotopic subspace 97° and a Krein space 97'. Moreover, if 2' = 20 [+] 2'" is another direct decomposition, then 2" is also a Krein space (Langer [9] ). Hint: Use the Definition 2.2 of a Krein space and Proposition 1.23.
Canonical projectors P± and canonical symmetry J
§3 1
Suppose a Krein space 0 is given, with a canonical decomposition
-. This decomposition generates two mutually complementary projectors P' and P- (P' + P- = I, the identity operator in .-W) mapping .W on to 0' and .,Y- respectively. Thus, in the notation of §2, for any x E 'W we have P±x = x±. The projectors P' and P- are called canonical projectors. Recalling the relation (2.5) we observe that P± are ortho-projectors, i.e., they .,Y = .-W' [+]
are orthogonal (self-adjoint) projection operators relative to the scalar product (2.2):
.W =,N,+ Q+ .) o = P'.
O+ P-'.'.
We now bring into consideration a linear operator J: --W formula
J= P' - P-,
.# defined by the (3.1)
§3 Canonical projectors P± and canonical symmetry J
19
and call it the canonical symmetry of the Krein space. The justification for this name is that, firstly, the operator J is defined by the canonical decomposition (2.4), and secondly, that it is a bounded symmetric (self-adjoint) operator in, since it is the difference between two self-adjoint operators P'
and P. 3.1
The canonical symmetry J has the following properties:
J*= J
J2=I(J-'=J) J-' = J*
(self-adjointness), (the property of being involutory) (the property of being unitary).
(3.2) (3.3)
(3.4)
The relation (3.2) was established above. Further, for x = x+ + x- E
we have, by (3.1), Jx = x+ - x-, and so J2x = J(x+ _X-)=X, +X- = x, i.e., J2 = I. Finally, (3.4) follows from (3.2) and (3.3).
2 We trace the action of the operator J on the subspaces+ and-. It follows from (3.4) that
3.2 M+ is an eigen-subspace of the operator J, corresponding to the eigenvalue k = 1. 3.3 j e- is an eigen-subspace of the operator J, corresponding to the eigenvalue X = - 1.
Since the operator J is simultaneously both self-adjoint and unitary (see 3.1), its whole spectrum lies on the intersection of the real axis and the unit circle, i.e., on the join of the points Xi = 1 and X2 = - 1. As we see from 3.2 and 3.3, it actually (provided only that ,Y+ ;e (8) and - 7 (81) contains both these two numbers, which are eigenvalues of the operator J: a(J) = up (J) = (- 1, 1). We recall that a(-) denotes the whole spectrum, and up(-) the point spectrum, of an operator. We notice further that the definition J(x+ + x-) = x+ - x- of the canonical symmetry operator J enables us to treat the result of its action as the `mirror reflection' of the space . in the subspace +.
The introduction of the operator J makes it possible to reunite in a new and more compact way the fundamental relations derived in §2 between the indefinite metric ([x, y]) and the Hilbert metric ((x, y)) in a Krein space, 3
namely: 3.4
(x, Y) = [Jx,Y],
[x,YA =(Jx,Y)
(x,YEJY)
(3.5)
1 The Geometry of Spaces with an Indefinite Metric
20
The first of these relations is the equivalent of formula (2.2), and the second is the equivalent of formula (2.6). It follows from 3.4 that, firstly, 11x11'= [Jx,x], and secondly that: 3.5
[x,x]=(Jx,x)
(xE.0),
(3.6)
The form [x, y] is a bounded (and therefore continuous over all the
variables) Hermitian sequilinear functional on ' wx
with the bound (since
11 JlI=1)
(x,yE.W).
[x,Yll<11ill lIxliiiyll=IIxllIIyII
(3.7)
The Proposition 3.5 enables an important deduction to be made about the closures (with respect to the norm 11 x 11) of semi-definite lineals of a Krein space:
3.6 If 2' C
2C
+ (? - ), then also 2' C R
If 91 C .- °, then also
°.
From this follows, in turn, 3.7 In a Krein space every maximal lineal 2' (C .? +) (or 2' (C-40-), (C o° 0)) is closed, i.e., it is a subspace: 2' = Se We note that for maximal definite lineals a similar assertion is certainly not true (see Example 4.12 below). 4
To the properties of the operator J of being symmetric x, y E .
Y) = (x, Y ),
and of being unitary, and therefore isometric
x,yE.W, which reflect its properties relative to the scalar product (x, y), can be added its entirely analogous properties relative to the indefinite metric [x, y]:
[ x, Y] = [x, Y], [ x, Y] = [x, Y]
(x, Y E 0),
(3.8)
which follow from (3.5): [ Jx, Y] = (x, Y) = (Y, x) = [ Y, x]
[Jz, Y] =(x,
(x, Y E .,Y);
(x,YE.#). These properties have been named the J-symmetry and the J-isometricity respectively of the operator J; they are discussed in detail later in Chapter 2, [x, Y]
Sections 3.5. As regards the indefinite metric [x, y] itself, in view of the second formula in (3.5) it is also known as the J-metric, and a Krein space is called a (Hilbert) space with a J-metric or, shortly, a J-space.
§3 Canonical projectors P± and canonical symmetry J
21
For the same reason the Q-form Q(x, y) = [x, y] = (Jx, y) is called the J -form, and correspondingly we speak of J-orthogonality (x [1] y) of vectors x, y E . , of the J-orthogonal complement -It fly to a set ,Jf (C.'), and so on.
We shall adhere to this latest terminology throughout all the rest of the exposition unless otherwise stipulated (see, e.g. §9, para. 4). In the theory of linear operators in Krein spaces (see Chapter 2 and later) it is often convenient to regard. Was ' a Hilbert space in which an indefinite metric 5
is defined not by a canonical decomposition . =+ @,k- given in advance, i.e., by canonical orthoprojectors P± (P+ + P- = I) (see §2.4), but by some preassigned symmetry operator J. For this purpose it is expedient to modify the definition (as will be seen in a moment, to an equivalent form) of the symmetry operator J and, at the same time, also of a canonical decomposition and of canonical projectors by putting as a basis not the form (3.1) but the
relations (3.2)-(3.4), any two of which imply the third (see the Proof of Proposition 3.1).
Definition 3.8: Any linear operator J: .Y - W' which is simultaneously unitary and self-adjoint:
J-'=J*=J is called a canonical symmetry in the Hilbert space M. A canonical symmetry J immediately generates orthogonal canonical orthoprojectors P± according to the formulae
P+='(I+ J),
P- =2'(I- J),
and a canonical decomposition
and also the J-metric
[x,yl =(Jx,y)
(x,yE (x,yEA),
where (x, y) is the original scalar product in ,Y, and the subspaces
.
±,
relative to the J-metric, are definite (of different `signs') and J-orthogonal. We
mention that here the `limiting' particular cases where J= ±I, i.e., either - = [e] or W+ = (e}, and the J-metric [x, y] is identical with either (x, y) or - (x, y), are not excluded. Both in the general theory and (in particular) in applications the canonical symmetry operator J may be defined in the Hilbert space in very different ways. We illustrate this fact here by two examples (the reader will encounter others in the later sections and chapters). Example 3.9: We consider a Hilbert space .W' = .
,
O+ ..*2 which is the
orthogonal sum of two copies of a Hilbert space ., (= . , = ,Y2), and we
22
1 The Geometry of Spaces with an Indefinite Metric
define in .
an operator J in the 'operator-matrix' form J=
G
(I- G2)"2VI
V*(I- G2)ti2
- V*GV
(3.9)
where G is a bounded self-adjoint operator, and I is the identity operator in ./', with 0 5 G <, I, and V is a semi-unitary operator in W' mapping .( on to V.Y( D [Ker(I- G2)]1 (actual examples of such a pair G, V will be met later in Example 1.4 in Chapter 3 and elsewhere), or V is an arbitrary unitary operator (V*V = VV* = I).
In the matrix form (3.9) the operators (I- G2)I/2V and V*(I- G2)1/2 appearing on the secondary diagonal are to be understood as operating, the first from 2 into y'i, and the second from 4', into '2. Thus, strictly speaking, they should be written in the form (I - G2) 1/2 Vj12 and j21 V* (I - G2) "2, where j12 and j21 are the operators of `canonical imbedd-
ing' (of identification of elements) of 2 into
1
and oft into W2
respectively. In order not to complicate the notation we shall not do this either now or later, hoping that no misunderstanding will arise in the reader's mind. Clearly j* = J, and so it remains to verify that J2 = I (the identity operator in ). But it is easy to see this by squaring the matrix (3.9) and noting that
(I- G2)
C [Ker(I- G2)]1 and GV.
C [Ker(I-
G2)]1.
The most important cases of the canonical symmetry operator (3.9), those most often used in applications are the `extreme cases':
G=I,
V=I: J=
G=0,
V=I: J=
(3.1Oa) 0
I
I
0
(3.10b)
_ representation We note that only for the first of these is the _original = M, ( M2 also a canonical decomposition Se- = + [+] Se- - (.ie- + = .
.H1, .
- = W2). In other cases the canonical decomposition has to be con-
structed afresh (see Theorem 8.17 below).
2 (., =
2 =), but this time Example 3.10: Again let if- =.,Y, O+ .Ye itself is a Hibert space with a J-metric, i.e., it is a Krein space. We introduce in . ie- the canonical symmetry operator J=
0
eJ
eJ
0
(3.11)
Here again the self-adjointness of j in. is obvious, and the formula j2 = h an be verified immediately. In the particular case when J = I, c = 1 the formula (3.11) turns into (3.10b). Example 3.11: Consider the Hilbert space _W =12 of all infinite (in both (x, y E 12) with square-summable directions) sequences x = y = (n;);
§3 Canonical projectors P± and canonical symmetry J
moduli, and with a scalar product ( , ) and an infinite form
23 [
,
given
]
respectively by the formulae (cf. Example 1.33)
(x, y) = E
Aj,
[x, y] = z
'rl -j- 1.
Here the operator J(j;); _- = [E-j- 1); _m is an involution and J* = J.
Exercises and problems 1
In Exercise 3 of §2 construct the canonical symmetry operator J from the given canonical decomposition.
2
Construct the canonical symmetry operator J for the space in §2, Exercise 4.
3
Do the same for the spares in §2, Exercises 5 and 6.
4
In a Krein space the lineals 2 and J/ are always skewly linked: Y # J. (see § 1, para 10) ([V]) (cf. Exercise 6b below).
5
Find the canonical decompositions of the Krein spaces in Examples 3.10 and 3.11 corresponding to the canonic symmetry operator J indicated therein. The concept of a canonical symmetry operator makes sense also in any space t with a Hermitian form Q(x, y) = [x, y] (x, y E 4T) if . admits a decomposition . =.:+ [+]. -, which, clearly, defines projectors P± from ? on to . respectively
6
(P' + P- = I). Verify that (cf. Scheibe [1]):
a) J = P+ - P- is an involution: Jz = I, and a Q-symmetry: [ix, y] = [x, Jy] (x, y E fl, and is a Q-isometry: [ Jx, Jy] = (x, y] (x, y E j); b) For any lineal 2' (C. f) the lineals 9 and J2' are skewly linked (2' # J2') and are Q-isometrically isomorphic (see § 1, para. 3); here a Q-metric is induced in 99 and Jc
from .1;
=Jell(SPC ); in particular,
c)
if 9 is maximal in the class of non-negative (positive, neutral, non-positive, or negative) lineals, then J2' is also maximal in the same class; a similar assertion holds
for non-degenerate lineals, and if 2'' [11 = 2', then (J2')[1] [11 = JY; d) for a neutral 2' the lineal 2' + J2 admits the canonical decomposition ' + J2' = P+ 2' [+] P- SC, and moreover P' 2' is Q-isometrically isomorphic to the anti-space for P-2' (see §1, para. 3). 7
I f ,Y = 99 Q+ 21 is the orthogonal sum of two subspaces of a Hilbert space .
, then in
order that the matrix 11
J12
1z
zz
11
with bounded elements should give (relative to the given decomposition of Y) a canonical symmetry operator in 0, it is necessary and sufficient that a) 11 = Jll, J22 = J22;
b) Ji1 + 12J z = IV';
c) Jz2 + Jiz 1z = IT- ;
d) J11 J12 + J12 J22 = 0.
Hint: Verify the condition Jz = I.
24
§4
1 The Geometry of Spaces with an Indefinite Metric
Semi-definite and definite lineals and subspaces in a Krein space
Let W = W+ [ Q+ ] .- be a canonical decomposition of a Krein space, and let P+- be the corresponding canonical projectors. The symbol [(@] is used here (for the first time) as a brief remainder of the 'two-fold orthogonality' (the usual one and also in the J-metric) of the canonical decomposition we have fixed on. We consider an arbitrary semi-definite (for definiteness, a 1
non-negative) lineal 2' (C.') and we study the mapping P+: Speaking more accurately, we actually study here the restriction P+ 12' of the operator P+ on to 2', and the image of 2' in this mapping is the lineal P+Y
Theorem 4.1: The mapping P+ 12': 2' -+ P+2' of a non-negative lineal 9.9 (C.V) is a linear homeomorphism, i.e., it is bijective, continuous, and the
inverse mapping (P+ 12)-': P+2- 2' is also continuous. The linearity and continuity of the ortho-projector P+, and also of its restriction P+ 12', are obvious. The bijectivity follows from Lemma 1.27. Further, for x E . we have (see 2.6) II x+ 11 > 11 x 11, and so II(P+I2')x112=l1x+11>2I(IIx+112+11x
(xEY).
(4.1)
Putting (P+ 12')x = y, we have x = (P+ 12) ' y, and the relation (4.1) gives II (P+ 12'y'y 112
211 v 12
(y E P+E), i.e. II (P+ 12')-' II < J.
Corollary 4.2: Under the conditions of Theorem 4.1 the lineals 2' and P+2' are either both closed or both are not closed. Corollary
Se (C,f = < dim .+.
4.3: + [+]
dimension of any non-negative subspace does not exceed the dimension of .Y+: dim .'
The )
Remark 4.4: By virtue of Lemma 1.27 the bijectivity of the mapping (P+ I SP): 2' P+Y in Theorem 4.1 holds for any linear space ; with a Q-form [x, y] (see §1, para. 1) which admits decomposition into a direct sum (not necessarily even a Q-orthogonal sum!) .4'= . + + - of a positive lineal (.:W'+) and a negative lineal (.: - ), if P± are the projectors corresponding to this
decomposition (P+ + P- = I.y, the identity operator in J r). So in this case Corollary 4.3 still holds if in the inequality dim 2 < dim 2+ we take the symbol dim to be not the Hilbert dimension but the linear dimension, i.e. the cardinality of the algebraic basis. Finally, even if Jr .is degenerate, if it still admits decomposition into the direct sum . = .t ° +;+ +;- of an isotropic lineal ;T0, a positive lineal ;+, and a negative lineal .: - (see 1.24), then similar arguments show that the linear dimension of any positive .' (C.Jr) does not exceed the linear dimension of +.
§4 Semi-definite and definite lineals and subspaces in a Krein space
25
In conclusion we point out that we have, only for the sake of definiteness, been dicussing non-negative and positive lineals. The reader will easily be able to formulate and prove analogues of all these assertions for non-positive and
negative lineals. But, actually, repetition of the proofs is unnecessary here because a simple transition from the space . (respectively .) to the anti-space (see Definition 1.8) changes the roles of non-negative (positive) and nonpositive (negative) lineals, in particular of ( +) and J (. ) and correspondingly of the projectors P+ and P-. The results of paragraph 1 open the way to the establishment of criteria for maximality of semi-definite lineals in a Krein space. 2
Theorem 4.5: In order that - (C,41+) (2' (C,01-)) in the Krein space .,Y =+ [ O+ ],W- should be a maximal non-negative lineal (maximal nonpositive lineal) it is necessary and sufficient that P+ 2 = W+ (P-2' = W-). El
We carry out the proof for a non-negative Y. Necessity: Let 22 be the maximal lineal from 40 +. Then (see 3.5) 99 is closed,
i.e., it is a subspace, and therefore (see Corollary 4.2) P+Z ' (C, +) is also a subspace. Let us assume that P+22 ;6 A0+. Since onk+ the metrics (x, y) and [x, y] are identical (see (2.2)), we can find a vector (9;4)zo E 99+' = .,Y+OP+Y22
which is simultaneously orthogonal and J-orthogonal to P+2'. 9 is called the deficiency subspace for Y. Thus zo 1 P+2' and zo [1] P+2'. At the same time zo [1] 2', because for any x E 2' we have [zo, x] = [zo, x+ + x-] = [zo, x+] = (zo, x+) = (zo, Px+) = 0.
We note further that zo !f 22, for otherwise it would be isotropic for .' and therefore a neutral vector, but zo is positive (0
envelope Lint9,zo] =
C P+, and also
.
Zo E
+ ). But then the linear 22, which con-
3 2' and 22
tradicts the maximal non-negativity of Y. Sufficiency was established earlier in Corollary 1.28.
Corollary 4.6: If 9 is a maximal non-negative (maximal non-positive) subspace of the Krein space W =,W+ [ @].,Y-, then dim 22 = dim ,Y+ (dim 2' = dim ..- ). El
3
This follows immediately from Theorems 4.1 and 4.5. The situation is more complicated with maximal positive (maximal nega-
tive) lineals, which, as we shall see below (Example 4.12), can also be non-closed. We start, however, from a simple fact: Theorem 4.7:
In order that 91' C
++ U 101 (2'
C --- U t9)) should be a
1 The Geometry of Spaces with an Indefinite Metric
26
maximal positive (maximal negative) lineal, it is necessary, and in the case
when .' is closed, it is also sufficient, that P91 =+ (P- = ,Y- ). Necessity. Suppose for definiteness that 2' is a maximal positive lineal. If +, then, again choosing an arbitrary vector zo ;d 0 in the
P'
X e Q PP Y7, we discover, exactly as in the proof of Theorem 4.5, that 2' = Lin (2', zo) is positive, and also Y D 2' and 2' ;6 2', deficiency space C A y,
which is impossible.
_
Sufficiency: Let P _ W'. If the positive 2' is closed (91 =2'), then by Corollary 4.2 P2' _ = j+ is also closed. By Theorem 4.5 2' is a maximal non-negative subspace and by the same token a maximal positive subspace.
In the course of proving the `sufficiency part' of Theorem 4.7 we have incidentally established.
Corollary4.8: A maximal positive lineal Yin the case when it is closed is also a maximal non-negative lineal.
Remark 4.9: closed, the sufficient for
In the case when 2'C.#++ U (01 (2' C .40- - U (B)) is not
condition P =+ (P =M-))
the maximality of .',
is certainly not because it is also possible that
2'C ,-++ U (9) (k c 0--U 101). As the simplest example in the case, say, of an infinite dimensional i+ we may take any lineal 2' (?! ,W+) which is dense in
In view of Remark 4.9 the criteria (i.e., the necessary and sufficient conditions) for maximality of definite lineals have to be sought in other terms. Theorem 4.10: In order that in the Krein space =+ [ @],W- the lineal 2' C _°++ U (0) (2' C .:-- U (0)) should be a maximal positive (maximal negative) lineal, it is necessary and sufficient that the following two conditions hold:
(1) The closure 2' of the lineal 2' is a maximal non-negative (non-positive) subspace. _ _ _ where (2')° is the isotropic lineal for 2' (not excluding the (II) = 9' + possibility (9')° = (0)). Necessity. Suppose, for definiteness, that 2' is a maximal positive lineal. virtue of Theorem 4.7 we have _P Y =+ and a fortiori P+2'= +, and therefore (Theorem 4.5) JR is a_ maximal non-negative subspace; so (I) has been proved. Further, 9' C 2', (2')° C 2', and in addition Pfl (2')° = (0), because 2' is positive and (2')° is neutral. But then 2'= 2'+ (.9')° + 2',, where 21 is positive or equal to (0). But since the positive lineal 2' + 2' > 2', so, by virtue of the maximality of 2', we have 2i = (0), i.e. (II) has been proved. Sufficiency: Let the conditions (I)-(II) hold for a positive Y. We consider By
§4 Semi-definite and definite lineals and subspaces in a Krein space 27
the maximal positive lineal .max (3 9); it exists by virtue of the maximality principle (Theorem 1.19). Then, on the one hand, 9max 3 2 and because of (I) 9max = 9, but because of (II) Jmax =
_ Y + (Y)°.
(4.2)
Now suppose there is a Z E Wmax\9(C 9max C 2max) Then, in accordance with (4.2), we have z = x + xo (x E 9, xo E (JP) 0) and Xo = Z - x E 9max n (2')°, i.e. xo = 0 and z = X E 9, contrary to the choice of z. Thus, 9max = Y. Before giving an example of a non-closed maximal positive lineal, we recall
one lemma of a general character, which we shall have occasion to use more than once later. Lemma 4.11: Let '1' be an infinite-dimensional normed linear space and let e E 4 e ;w-I 0. Then there is a linear 9 (C A') such that
k=,il
/t
Lintel.
(4.3)
Without loss of generality we shall suppose that II e II = 1. As is well known, there is an unbounded (discontinuous) linear functional gyp: A'- C such that p(e) = 1. Let 9 = Ker (p, i.e., the set of zeros (the kernel) of the functional (p. Then (see, for example, [XIII]) 9 = , A At the same time for any x E ./V, having chosen X = po(x), for xo = x - Xe we have gp(xo) = fi(x) - X = 0,
i.e. xoE9andx=xo+Xe. Example 4.12: We consider a Krein space .
, where dim .,Y' = ao and (e,+ )a E A is a complete orthonormal system in .W+, but
dim Y- = 1, i.e. W- = Lin(e- l,
11 a
= .O+ [
].
11 = 1. Thus
[e,+,evl =(e,',eo)Saa, [e.+,e-]=(eq,e-)=0,
[e,e]=-Ilel12=-1,
(«,0EA), (4.4)
where A is some (infinite) set of indices. We fix in A a certain index ao and consider the subspace 9i= C Lin (e o + e-; ea J. ,E A, a 0 ao By (4.4) this subspace is non-negative, and moreover it is maximal non-negative, because P+9 = C Lin (ea la E A =,,Y+. The vector e = e o + e- (;60) is neutral, and so
Lin (el = k° is an isotropic subspace for 9 (see 1.17 and the analogue of Corollary 4.3 for non-positive subspaces). We now apply Lemma 4.11 and express in the form 9 = 9 + Lin (el = 9 + 9°, where L = 9. Further, 2' is
positive, 2'(= I) is a maximal non-negative subspace, (9)° = §'° and r = 9 + (9)°. By Theorem 4.10 9 is a maximal positive lineal (and is, moreover, non-closed) (see Exercise 2 below).
4 We now consider another particular case of semi-definite lineals in a Krein
space, that of neutral lineals 9 (C.?°). Since such an 9 is simultaneously
1 The Geometry of Spaces with an Indefinite Metric
28
non-negative and non-positive, both the mappings
': Y -- P+T (C ,Y+) and P- I ': 99 - P V (C ,Y-)
P+ I
are linear homeomorphisms (Theorem 4.1). Theorem 4.13: In order that a lineal 91 (C.:e°) should be a maximal neutral subspace it is necessary and sufficient that one of the conditions
(I) P+Y=. +;
(11) P-F= Y-.
should hold. Necessity: Let 2 be a maximal lineal from jo° and therefore closed (see 3.5). If both the conditions (I) and (II) are infringed, then in the (non-zero) deficiency spaces 9v = . + O P-21 and CAv = . - O P-11 we choose vectors x+ and x- respectively (with II x+ II = II x II > 0). Then (see the proof of Necessity in Theorem 4.5) x+ [1]Y and x- [1]Y, and so x = x+ + x- [1]Y. Moreover, the vector x is neutral, because (see (2.7)) [x, x] = II x+ III_ II x 2 = 0. At the same time x %', because otherwise it would follow from the relation x+ 1 P+.T for instance (with x+ chosen in the same way) that II x+ 112 = (x+, x+) = (x+, p+x) = 0, contrary to the condition 1I
I
I x+ I I > 0.
It remains to consider the neutral lineal 9 = Lin(', x) 3 2
(9 .') to arrive at a contradiction with the maximal neutrality of Y.
Sufficiency: If, for example, condition (I) holds, then the neutral subspace T is, by Theorem 4.5, maximal non-negative, and therefore is also a maximal neutral lineal.
Corollary 4.14: Every maximal neutral subspace is either a maximal nonnegative subspace, or a maximal non-positive subspace, or is simultaneously both the one and the other. This follows from the conditions (I), (II) of Theorem 4.13, and Theorem 4.5.
Definition 4.1S: If a maximal neutral lineal 2' is simultaneously both maximal non-negative and maximal non-positive, then it is called a hypermaximal neutral linear. Remark 4.16: From the Definition 4.15 and Corollary 4.6 it follows that +-, and so maximal neutral subspaces can exist only in those Krein spaces .,Y _ .,Y+ [ S] .,Y- in which dim -At" = dim i .
dim .2= dim
To conclude this section we indicate other criteria for neutrality, maximal and hyper-maximal neutrality of a lineal I (C.,Y) which can be formulated 5
and proved without reference to a canonical decomposition .WP = .W+ [ + ] .,Y-,
i.e., in a form valid for arbitrary spaces .4 with an indefinite metric (see §1).
§4 Semi-definite and definite lineals and subspaces in a Krein space 29
4.17 For neutrality of 99 it is necessary and sufficient that 2' C 2''. If 2' is neutral, then all its non-zero vectors are isotropic (see 1.17), and therefore 2' C 2' . Conversely, it follows from this inclusion that the lineal Y,O = 2'n27[1J isotropic for 2' coincides with 91 itself, i.e., - is neutral. 4.18 In order that 2' (C .-_jO°) should be maximal in _0°, it is necessary and sufficient that 2' be semi-definite and 211J [ll = 99. If 2' is a maximal linear in a°, then by proposition 4.17 9,1 C Yr1I, and therefore cannot be indefinite, since otherwise in any decomposition I'J = 2' + 2, the lineal 2, would also be indefinite (since 2' [1] 2',) and so, by virtue of 1.9, it would contain neutral vectors not falling within 2', and this
would contradict the maximality of 9' (C.?°). Thus 2'[1] is semi-definite. Further, since from Proposition 1.17 we have that any neutral vector from is isotropic for 2'[1}, by virtue of its maximality 2' (C -0 °) must coincide with the isotropic lineal (2'W)0, i.e., 9' = Y[11 n2'[1l 1l1 But Y[1) [1] D 9' (see (1.11)), i.e., Y[1] [1] = Y. Conversely, let 2' C YI1), let 2,[1] be semi-definite, and Y(1] [] = Y. Let
' J 2', where
c ,e °, and there is an xo E 2'\ Y. By Proposition 1.17 the
vector xo is isotropic for 2,111
and a fortiori xo [1] .', i.e., xo E Y[1] . But because
is semi-definite the natural vector xo is also isotropic for 2''J, i.e.,
xo E 2,[1] n 211] [1] = Y[l] n 2', and therefore xo E 2', contrary to the choice of xo.
4.19 From hyper-maximal neutrality of 2' it is necessary and sufficient that 2' = Y(1]
From maximality neutrality of 2' alone it follows, by 4.18, that 2[l]
(3 99) is semi-definite, and from the hyper-maximal neutrality of 2' it follows,
by virtue of the Definition 4.15, that 2'= If now the then, by Proposition 4.17, 2' C Conversely, if 2'= i.e., lineal f D 2' is semi-definite, then 2' [1] 2', and therefore 2' C 991-I _ Y. For neutral subspaces 2' in a Krein space = W+ [ O+ ]. - Proposition 4.19 enables us to obtain the curious formula: If the subspace 2' C -0 °, then 2I' = 2' [+] 1 i' [+] 1'v , (4.5) where 1, = ± O P±2' are the deficiency subspaces for Y. The inclusion 2' [+] 1' i1 [+] y C 211 I is obvious (cf. Proposition 4.17 and the proof of Theorem 4.13). Conversely, let z E Yl1]. Having expressed = . ' [O] (1% 1011: -,Y' in the form , Y ± = P±2' [ Q + ] 9 , we have 4.20
where
' = P+2' [ Q+ ] P- 91 is obviously again a Krein space, and 2' (C h') is a
hyper-maximal neutral subspace in ,Y'. Now for x = x+ + x- (x± E .+-) we have x± = x1 + xf , where x; E P± 3 (C .,Y') and xf E 1I . But x = (x; + x,-) + (xz + xz) [1] ' and xz + xZ [1] 2'. Therefore (,N")) x; + x; [1] 2', and since 2' is hyper-maximal in . Y ' , so (cf. 4.19) x; + x; E Y.
1 The Geometry of Spaces with an Indefinite Metric
30
With formula (4.5) and the argument we have just given there is closely connected the following Proposition, which will be extremely useful later: 4.21
If under the conditions of Proposition 4.10 Y' is a maximal non-
negative (non-positive) subspace in the Krein space 9y [+] 9 i', then 2' + 91 is a maximal non-negative (maximal non-positive) subspace in k.
We restrict ourselves to the case 2' C .-+ and (see Theorem 4.5) P+91' = 9y-. Then P+(2P'
+2')=Lin(P+Y',P+2')=Lin(9i,,+2')=.Y+.
It remains to use Theorem 4.5.
Exercises and problems 1
If A, and 9'2 are maximal non-negative (maximal non-positive) lineals, and 2' = Lin (2'1, 2'2], then for an isotropic lineal 2'° the inclusion 2,0 C 2',12'2 holds
([V]) 2
3
Prove that a non-closed positive lineal 2'i (C.7Y) exists with P+21 = + which admits a proper extension into a positive, but again non-closed, lineal 2' ( 3 911). Hint: As 2' use, for example, the maximal positive lineal in Example 4.12, and then, having again applied Lemma 4.11, construct on 2, C T. We consider for 2', (C.r') its deficiency subspaces Vi, _.°± O PP M'i. Let 2' be the maximal lineal from + and 2, C Y. Then dim(2' Q 2',) = dim Ce v-, ([V] ).
4
Give examples of hyper-maximal neutral lineals 2' in the space .,t' from Example 3.11.
5
In the real, two-dimensional, Krein space (see Exercise 7 to §2) distinguish by shading in Fig. 1 the sets :e + and e°- (see (1.7)), and find positive, negative, and neutral lineals. Verify that they are all maximal (in their respective classes) and that the neutral lineals are even hyper-maximal.
6
Show that in a Krein space Proposition 4.19 and 4.20 (i.e., Formula (4.5)) are equivalent.
§5
Uniformly definite (regular) lineals and subspaces. Subspaces of the classes h'
In this section we shall mainly discuss certain subclasses of definite lineals 1 and subspaces in a Krein space Y = 0+ [ O+ ] . W-. In this connection the basic
concept by means of which these subclasses will be distinguished is the so-called intrinsic metric. Suppose, for definiteness, that the lineal 2 is positive. Since the form [x, y] is positive definite ([x, ] > 0 for all (0* ) x E 2'), it can be adopted as a new scalar product on 9"; by so doing, 2' is converted into, generally speaking, a pre-Hilbert space.
Definition 5.1: Without introducing new designations we shall call the
§5 Uniformly definite (regular) lineals and subspaces
31
restriction of the metric [x, y] on to 2' the intrinsic metric on 91, and the corresponding norm
Ixl i= [x,x]1/2
(XE2')
(5.1)
will be called the intrinsic norm on Y. From (3.7) it follows that the estimate
IxI'sIIxMM
(xE2')
(5.2)
holds for the intrinsic norm I x I r- on 2' introduced by formula (5.1), i.e., I x I r is less than the basic ('exterior') norm II x II on Y. Our first aim is to distinguish the case when these two norms are equivalent.
Definition 5.2: A positive lineal 2' is said to be uniformly positive if there is a
constant a > 0 such that
Ixly
aIIxII
(5.3)
(xE2e)
or, what is the same thing, [x, x] > a2 II X II2
(5.1)
(xE 2').
It is clear that (5.3) combined with (5.2) means that the norms I x , and II x II are equivalent on Y. Entirely similarly-that is, by the same requirement (5.3)-we introduce the definition of a uniformly negative lineal 2'; this time the intrinsic norm I x I v' on a negative lineal 2' is defined by the equality I xI Y'- I
[x,x]I1/2=(-[x,x])1/2
(XE2').
(5.5)
In this case the relation (5.4) becomes
-[x,x]>a2IIxII2
(xE2').
(5.6)
Uniformly positive and uniformly negative lineals 2' are combined under the general name uniformly definite lineals. Uniformly definite lineals have an obvious 'inheritance' property: 5.3 Every lineal which is contained in a uniformly positive (uniformly negative) lineal is itself uniformly positive (uniformly negative). The simplest examples of uniformly definite lineals are the components
of the canonical decomposition N1 = W [ O+ 1,W-, and also, by virtue of Proportion 5.3, all lineals which are contained in -W+ (.W- ). 5.4 All finite-dimensional definite lineals are uniformly definite. This follows from the fact that in a finite-dimensional space 2' all norms are equivalent. (including II - II and I 5.5 91 is uniformly definite if and only if its closure, the subspace 2', is uniformly definite.
1 The Geometry of Spaces with an Indefinite Metric
32
This follows immediately from 5.3 and the fact that the inequalities (5.4) and (5.6) continue to hold on passage from 2' to its closure 2' (see proposition 3.5).
2
The question of the connection between the closedness of a definite lineal
(C.) relative to the exterior norm and its completeness relative to the intrinsic norm I I r (intrinsic completeness) is of particular interest. When 2' is uniformly definite, i.e., when the norms II ' II and I ,,, are equivalent, it is I
clear that the two properties just mentioned either hold or do not hold simultaneously. This assertion admits a partial inversion:
5.6 A definite subspace 2' (=2') is uniformly definite if and only if it is intrinsically complete.
implies the Because Ye is complete, closure of 2' relative to completeness of 2' relative to the same norm. Now let 2' be also intrinsically II
complete. Since the norms II
' II and I
I
I v, are connected by the inequality (5.2),
these norms are equivalent, by a well-known Banach theorem. Conversely, uniform definiteness of .' implies the equivalence of the norms ' II and I I v, on Y. Therefore from the completeness of 9? in the norm II ' follows its intrinsic completeness. II
II
As regard non-closed definite lineals 2', however, the situation is more complicated; such a lineal 2' may be either complete or not complete relative to the intrinsic norm I I,-. We have actually already met the first of these two cases (the lineal 2' in Example 4.12) (see §9, Exercise 11 below). However, it is clear that in the case of intrinsic completeness such a (non-closed) lineal cannot be uniformly definite.
3
Later, in §6, inter alia we establish criteria (necessary and sufficient
conditions) for the uniform definiteness of closed lineals (of subspaces-see Proposition 6.11), and in §8 (Lemma 8.4) a method of describing all such subspaces is indicated. But here we return for a moment to the first examples of uniformly definite lineals-to the subspaces + and - in the canonical decomposition [(D],*-. In a certain sense these examples can be called models, as the next theorem indicates. Theorem 5.7: In order that a linear 2' (* (B)) of a Krein space i( should be a Krein space (relative to the original metric [x, y]) it is necessary and sufficient
to admit a J-orthogonal decomposition 2' = 2+ [+] 2- into two uniformly definite subspaces: a positive one 2+ and a negative one 2-. Here it is not excluded that 2+ {B) or 2- = (0). Necessity: The necessary condition follows from Definition 2.2 of a Krein
§5 Uniformly definite (regular) lineals and subspaces
33
space, from the Definitions (2.2) and (2.3) of the scalar product and norm II in a Krein space, and from the Definitions (5.4) and (5.6). Sufficiency follows from Proposition 5.6 and the Definition (2.2).
'
II
Corollary 5.8: If 91 is a neutral subspace of a Krein space, then the factor-space . Y _ 2Ill/2' is again a Krein space (cf. Exercise 10 in §2). Recalling formula (3.5) we have 2 X11 = 2'[+]
i [+] civ , where
i [+] V i-
is a Krein space by 5.3 and Theorem 5.7. It remains to apply 1.23. The concept of a uniformly definite subspace admits a certain generalization which will later play an important part in the theory of certain classes of linear operators in Krein spaces (see Chapter 3, §5). 4
Definition 5.9: A non-negative (non-positive) subspace 2 of a Krein space .*'
is called a subspace of class h+ (class h-) if it admits a decomposition .' = 2'o [+] _T+ (2' = 2'o [+] 2-) into a direct J-orthogonal sum of a finitedimensional isotropic subspace 2'° (dim 2'° < co) and a uniformly positive (uniformly negative) subspare 2+ (9?-). A caution here is necessary: the fact that a subspace 2' belongs, say, to the class h+ does not in any way imply that in every decomposition 2' = 2° [-+] 2' the positive component (the lineal 2'i) will be uniformly positive. Example 5.10: In Example 4.12, the subspace 2E h,, because it admits the which in = Lin(e o + e-) [+] C Lin (ea'] a E A,,,,* o, decomposition Lin (e o + e- } is a one-dimensional isotropic subspace, and (Lin (e.+}a E A, a # ao is, by 5.3, uniformly positive. On the other hand, the same 2 can be represented (see Example 4.12) in the
form 2 = (2')°[+]2, where 2' is by no means uniformly positive-for if it were, then by 5.5 its closure 2' would also be uniformly positive, but in Example 4.12 it was degenerate!
Exercises and problems 1
2
Any Krein .N'_ .w+ [(D],W' with infinite-dimensional .W± contains definite subspaces which are not uniformly definite (Ginzburg [4] ). In the case of separable -W± with orthonormalized bases (ek )k=' and (e k )7=1, consider, for example, the subspace 21 = C Lin I ek + (k/k + 1)ek) k=2; prove that it is positive but not uniformly positive (I. Iokhvidov [13]). Modify this example for negative subspaces and for non-separable -W±.
Let Y be an arbitrary lineal in a Krein space, and let , r, = Y'(1) n Y'1. Then (J, is a Krein space (see Ritsner [4] ). Hint. Prove that V y, = Cl i [+] 1 v, where J ry+ are the deficiency subspaces for Y (see
§4), and then apply 5.3 and Theorem 5.7.
34
1 The Geometry of Spaces with an Indefinite Metric
3
Prove that all maximal, uniformly definite lineals are closed, and obtain the maximality principle for them (cf. Theorem 1.19).
4
Every maximal uniformly definite lineal 2' is a maximal semi-definite lineal ([V] ). Hint: If 2' is positive, for example, then, assuming the contrary, consider its deficiency subspace 9 v' and prove that 2', = -'' [+] L( i is uniformly positive.
5
Let 2' (C.) ') be a definite lineal, and let yo E .w'. In order that the functional ,py = [x, yo] should be continuous relative to the intrinsic norm sufficient that
v, it is necessary and
m (yo) = inf [x - yo, x - yo] > - co (if 9' is positive), XEY'
M(yo) = sup [x, yo, x - yo] < oo
(is 2' is negative).
X E Y'
If these conditions hold for the (intrinsic) norm I rpo I Y' of the functional ,p, we have Pyo zY' = Y' = M(yo) - [yo, yo] respectively (I. Iakhvidov, [yo, yo] - m(yo) and see [VIII].
6 A definite lineal 2 ' (C ,Y) is said to be regular if for every y E . ' Y the linear functional py(x) = [x, y] (xE 2') is continuous relative to the intrinsic norm I IY-, and to be singular otherwise (I. Iokhvidov [7]; see also [VIII]. Prove that regularity of a definite lineal is equivalent to its being uniformly definite ([VIII)).
Hint: Extend by the Hahn-Banach theorem every linear functional which is continuous (relative to the external norm II ' II) on 2' into a functional continuous on the whole of .JY, and then apply F. Riesz's theorem, and thus show that the stock of linear functionals on 2' which are continuous relative to the norms II and is the same, from which the equivalence of these norms will follow. 7
Let the subspace 2' C .0°, and let J' = 2111/2' be the corresponding Krein space (see Corollary 5.8). Then maximal semi-definiteness in. ' Wof the subspace 2' is equivalent to the corresponding semi-definiteness in .J'Y of the lineal Lin [ x I x E z, I E 2).
Hint. Use (4.5), 1.23 and 4.21.
§6
Decomposability of lineals and subspaces of a Krein space. The Gram operator of a subspace. W-spaces and G-spaces
In §1, paragraph 9 we touched on, and in Theorem 1.34 considered in rather more detail, the question whether an arbitrary non-degenerate lineal 2' can be decomposed into the direct sum 2 = Y+ + 2- of positive and negative lineals 2'±. Here we shall narrow down the problem somewhat, since we shall discuss only lineals 2' in a Krein space Y and decompositions of the form 1
Y _ Y+ [+] Y-'
(6.1)
in which 2+ [1] 2-. Definition 6.1: If a J-orthogonal decomposition (6.1) exists for a lineal 2' (C, W), then SP is called a decomposable lineal; otherwise 99 is called an indecomposable lineal. However, even the transition to Krein spaces does not reduce the problems of
§6 Decomposability of lineals and subspaces of a Krein space
35
decomposability of non-degenerate lineals Y (non-degeneracy is a necessary condition, see 1.24). As an example we consider in the Krein space .0 of Exercise 3 to §2 the lineal V' consisting of all those functions p E .s which, on the interval [ - 1, 0] coincide with polynomials. It is clear that Y is dense in .t, and therefore it is a non-degenerate lineal. On the other hand, ' _ Y, 4- .Y'2, where Y, consists of all those functions tip E c which are equal to 0 on [ - 1, 0], and Y2 consists of all those ,p E .Y' which are equal to 0 on [0, 1]. Y, and Y2 are neutral (see the metric (1.20)) and by construction they are not isomorphic (the linear dimension of 21, is greater than that of 22). Hence (see Theorem 1.34) follows the impossibility in general of representing Y in the form' = Y+ + 91- where £t are definite, and a fortiori the indecomposability of Y.
At the same time, the transition to Krein spaces enables us to solve completely (and moreover in the positive sense) the problem of the decomposability of its closed lineals, i.e., of subspaces. We start from an elementary fact: 2
6.2 An isotropic lineal Y° of any subspace Yin a Krein space is closed, i.e., it is an isotropic subspace. This follows immediately from the fact that, for any x E ', the passage to the limit as n co in the equality [x°, x] = 0, which holds for any sequence of vectors
isotropic in c, x°
x° (E2) gives, by 3.5, [x°, x] = 0, i.e., the vector x° is
isotropic in Y (if x° ;4 0). Later is this section (and in a number of other questions) the concept of the Gram operator of a given subspace will play a leading part. Let 9 be a subspace
of the Krein space W = W+ [ (B ]. - with the canonical symmetry operator J = P+ - P- and the J-metric [x, y] = (Jx, y) (see §3, paragraph 1). We consider the restriction of this J-metric on to ': [x, y] y-, i.e., simply the form [x, y] with the arguments x, y traversing only Y. Since 2, like the whole of
.W, is a Hilbert space with the scalar product (x, y) (x, y E Y) and the norm 11 x 11 = (x, x)"2 (x E 21), and [x, y]
is a sesquilinear (and moreover a bounded
(see 3.5)) functional in 3' (or, more precisely, in 91 x Y), there is a unique bounded self-adjoint operator
Gr, operating in
([x,y]'r=) [x,y]=(Grx,y)
' such that
(x,yE t)
Let P, be the (Hilbert) orthoprojector on to T. Then for any x, y E T
(G,,x,y)= [x,y]
Pvy) = (PvJx,y),
and so G, = P-, (J I M'), where J 19' is, as usual, the restriction of J on to T. Definition 6.3:
the subspace P.
The linear operator G, = P,J I Y' is called the Gram operator of
1 The Geometry of Spaces with an Indefinite Metric
36
Noting that 11 G,, 11 = 11 P,-(J 12')11 < 1, we now use the spectral decomposition of the bounded self-adjoint operator G,,, in Y: 3
G,
X dE,(X),
(6.2)
where E,(X) is the spectral function (resolution of the identity) of the operator G,, and the integral in (6.2) is understood as the limit of the corresponding integral sums in the uniform operator topology (i.e., with respect to the operator norm; see e.g. [XXIII] ).
Suppose, for definiteness, that the spectral function E,(X) is strongly continuous on the right, i.e. E,(X + 0) = E1(X ), where E,(X + 0) = s - lim,,1 x E,,(µ) (strong limit). We bring into consideration three orthoprojectors in the Hilbert space 2':
C
dE, (X) = E,(- 0),
P11 =
Pi =
1
1'° = E,(0) - Ev( - 0),
dE,(X) = I, - E, (0),
(6.3)
J 0
where the symbol J=°, means that in setting up the corresponding integral
sums the value of E,-(X) at zero is here taken to be equal to E,(-0) (cf. [XXIII] ).
It follows from the properties of the spectral function that the three orthoprojectors defined in (6.3) are pairwise orthogonal, and, by their definition, i
Pi- + P°- + Pi =
dE,(X) = I, ,
i.e., they generate an orthogonal (in the Hilbert metric) decomposition of 2':
Y= 2'- Q Y0 0+ Y+
(2+ = Pry', 2° =
(6.4)
Theorem 6.4: In the decomposition (6.4) 2- and Y+ are pairwise J-orthogonal negative and positive subspaces respectively, and 2° is the isotropic subspace for T.
It should be made clear at once that, even when 2 ;4 (0), any one (or even El any two) of the subspaces on the right-hand side of the decomposition (6.4) may reduce to (0); we shall not mention this fact again in future. We start by proving that 2° is an isotropic subspace for Y. By definition 1'0 = Po 2' = (E, (0) - E,( - 0))2', i.e., 9o is the eigen-subspace of the operator corresponding to the eigenvalue X = 0 (more shortly, M'0 = Ker G,), and this is equivalent to 210 being isotropic in 91. For, if xo E 2' the relation xo [1] 2 is equivalent to the fact that, for all x E SC ([xo, x] =) (G,xo, x) = 0,
i.e., G,xo = 0.
§6 Decomposability of lineals and subspaces of a Krein space
37
We now prove that Y- is non-positive. If (B ;d )x E Y- = PIY, then, in accordance with (6.1)-(6.3) J
J
X d(E,(X)x,E,(-0)x)=
_
X d(E,(X)x, Pi'x)
Y d(E,(X)x, x) =
[x, x] = (Grx,x) = f
i
X d(E'(-0)Er(X)x,x).
1
J
(6.5)
But
((e,(X)x, x)
when X < 0,
(Ev(-0)E,(X)x, x) = t(E1{ - 0)x, x) = const when X 3 0. Returning to (6.5) we obtain [x, x] =
1
X d(E,(-0)E,(X)x, x) = f _ o X d(E,<X)x, x) < 0, (6.6)
J
J
i
since - 1 < X < 0, and (Er(9?)x, x) is a non-decreasing function. By means of similar calculations the reader will be able without difficulty to show that Y+ is non-negative and 27 [1] _T+. From what has been proved it follows in turn that 27 and Y+ are definite, because a neutral vector xo contained in either of them would be an isotropic
vector for this subspace (when x ;4 0) and therefore, in view of the Jorthogonality S7O [1] 2+ and 91+ [1] 2- already proved it would be isotropic Y+. But then xo E Y° and by (6.4) xo = 0. for the whole of _ 97 [ (@ ] Y° [ O+ ] We point out that in the course of proving Theorem 6.4 we have on the way established the proposition.
6.5 2 is non-degenerate if and only if Ker Gr-= [6). This follows from the fact that 991° = Ker G,.
4
By analogy with Definition 2.1 and generalizing it slightly we adopt the
Definition 6.6. If 91 is a subspace of a Krein space, then any representation of
2 in the form of a J-orthogonal direct sum 9, _ 2° [+] 2+ [4-] 2-
(6.7)
of three subspaces, of which ?O is isotropic for SP and 2P+ (2-) are positive (negative), is called a canonical decomposition of Y. The existence of such a decomposition (see (6.4)) for any subspace 2 is established in Theorem 6.4. Moreover, it is clear that the (isotropic) component 9P° of a canonical decomposition is defined uniquely (see 6.5). As regards the components 22±, their construction in Theorem 6.4 was carried out
1 The Geometry of Spaces with an Indefinite Metric
38
in a special way by means of the spectral decomposition of the Gram operator
G,, i.e., it was based on the scalar product (x, y). But the latter, as is well known, was defined (see (2.2)) by a fixed choice of the canonical decompo(or, equivalently, by the choice of the canonical sition *=.,Y' symmetry operator J; see §3, paragraph 5). Hence it is already clear a priori that other canonical decompositions for 2' are possible, for example, Y = 9'0 [+] Sei [+] yi ,
(6.8)
in which, generally speaking, 2't * 2' and 2'i ;4 2' (see, e.g., Theorem 8.17 below). So the next theorem has all the more interest: Theorem 6.7: (Law of inertia). Whatever the canonical decomposition (6.8) of the subspace 2' may be, it is always the case that
dim 2'1+ = dim q",
dim 2'i = dim 2-,
(6.9)
where 2' and 9'- are understood to be the corresponding components of any other canonical decomposition (6.7) of the same subspace Y. Starting from the canonical decomposition (6.7), say, we use the deduction made earlier at the end of Remark 4.4, by virtue of which we have that,
for the linear dimensions, and also in the present case for the Hilbert dimensions, Y+,
dim 21- < dim Y-.
But in this argument the roles of the decompositions (6.7) and (6.8) can be interchanged; hence (6.9) follows. So, let a certain canonical decomposition (6.7) of the subspace 2' be fixed. Definition 6.8: The trio (a y-, v,, wof cardinal numbers
7r, = dim Y+,
v,, = dim Y-,
w y, = dim 2'°
is called the inertia index of the subspace 2' and it is denoted by In Y. It follows from Theorem 6.7 that In 2' is an invariant of the subspace 99 relative to all its possible canonical decompositions (6.7).
The concept of the inertia index makes it possible to introduce for an arbitrary subspace 2' of a Krein space one further important characteristic, which will in particular come into the foreground in §9.
Definition 6.9: Let 2' be a subspace of a Krein space W' and let In 2' = (a,, vv-, w,-) be its inertia index. Then the cardinal number x r =
mina,-, v,.) is called the rank of indefiniteness of the subspace Y. In particular, the rank of indefiniteness of the whole space W = . denoted simply by x, i.e., x = min(dim x', dim x - ).
' [ Q+ ] .
is
§6 Decomposability of lineals and subspaces of a Krein space 5
From the results of §6, para. 3
(cf.
39
(6.1) and (6.2)) this obvious
proposition follows:
6.10 A subspace 2' of a Krein space is non-negative (non-positive) if and only if its Gram operator G f is non-negative (non-positive), i.e., (G,x, x) > 0 (<0) for all x E Y. A subspace 2' is positive (negative) if and only if (G rx, x) > 0
(< 0) for all (0 ) x E Y. We also obtain without difficulty a criterion for uniform definiteness of a subspace 2' in terms of its Gram operator Gv: 6.11 In order that a subspace 2' should be uniformly positive (uniformly negative) it is necessary and sufficient that its Gram operator G r should, for some a > 0, satisfy the inequality Gv, >, a21', (G,- < -a2I i).
Since [x, y] = (Gvx, y) (x, y E 2'), the inequality Gv- > azl, (
a2(x, x) (< -a2(x, x)) (xE 2') for the subspace 2' to be uniformly positive (uniformly negative) (see (5.4)). Corollary 6.12: For a subspace 2' of a Krein space ,Y to be itself a Krein space relative to the J-metric [x, y] it is sufficient that the point X = 0 should be a regular point for its Gram operator G,.,,: 0 E p (G v).
The condition 0 E p(Gv) is equivalent to the existence of an a > 0 such that X dEv(X)+
Gv-= -I
X dE-r(X) al
(the so-called `spectral gap' (-a2, a2) at zero). But then 2'° _ (0], and in the canonical decomposition given by Theorem 6.4, 91 = 2' [ @] Y-, the conditions of Proposition 6.11 hold for the Gram operator Gv = G,. 12±, and so the 2± are uniformly definite. It only remains to apply Theorem 5.7. Later (see Theorem 7.16) it becomes clear that the condition 0 E p(Gv) is not only sufficient, but also necessary, for 2' to be a Krein space.
6 We became acquainted above with the concept of the Gram operator G, of an arbitrary subspace 2' of a Krein space,-W. Starting from this concept we can consider a space more general than a Krein space, namely, a Hilbert space with an indefinite metric. Let be a Hilbert space with a scalar product (x, y) (x, y E .W') and norm X 11 = (x, x)"2 (x E ), and let W be an arbitrary bounded selfadjoint operator (W* = W) given on .,Y. Then the Hermitian sesquilinear form [x, y] = (Wx, y) defines in an indefinite metric which we shall call the
W-metric, and we shall call the space N1 itself with the W-metric a W-space. W is called the Grain operator of the space W.
1 The Geometry of Spaces with an Indefinite Metric
40
It is clear that all the results of §1 hold for a W-space (for Q(x, y) = (Wx, y) = [x, y]). However, it is also possible to assert that many (though by no means all) of the facts proved for a Krein space remain true for W-spaces. The most important of them is the inequality
I[x,y]IsII K'II IIxII II.II
(6.10)
which establishes the continuity relative to the norm 11 11 of the Hermitian -
sesquilinear functional [x, y] (the W-metric) over the set of variables (cf. Proposition 3.5). Further, it is easy to see that the non-degeneracy of a W-space W (or the non-degeneracy of the W-metric) is equivalent to the condition Ker W = (B), or, what is the same thing, 0 ¢ up( W) (cf. 6.5). In this particular case we shall denote the Gram operator W by the letter G and speak of the G-metric and the
G-space. It is not difficult to understand that if W is a W-space then the factor-space = .W'/(Ker W) with the induced W -metric (see 1.23) is a G-space.
A G-metric may be either regular (0¢ a(G) o 0 E p(G)) or singular (0 E a(G)).
6.13 A Hilbert space N' with a regular G-metric is a Krein space. If .NP is a W-space, then .
=(Ker W) is a Krein space if and only if
I w =dl w.
The first assertion was established, essentially, in Corollary 6.12. The second assertion follows from the decomposition le =.w [ @I Ker W and Proportion 1.23, by virtue of which W is W-isometrically isomorphic to Rw, and therefore the condition 4w = R w is equivalent to the regularity of the G-metric (i.e., the W-metric) in . . We note that a singular G-space too can be converted into a certain Krein space by a completion method. The precise result is formulated thus: L7
6.14 Let .W be a singular G-space, let mI < G < MI, let G = J`,,-o X dEx (Ea+o = E)) be the spectral decomposition of its Gram operator G, and let
P- = J,;,o_o dEa and P+ = Jo dEx be the orthoprojectors defined by it (P+ + P- = I. cf. (6.3)). For any x E . in .0 a new scalar product.
we write x' = P±x and we introduce
(X, y) = [x+, y+] - [x-, y-] and a corresponding norm II x 11
(x,
y E.W)
= (x, x )112 (x E .0) . Then the form [x, y] can be continuously extended (i.e. continuously relative to the norm 11 - II -) on to the completion .-Te of the space 'Y relative to this norm; . is a Krein space relative to the extended form [x, y]. 7 We note that the projectors P±, as also the Gram operator G, are bounded relative to the original norm; hence for the new norm we obtain the estimate
§6 Decomposability of lineals and subspaces of a Krein space 5 c II x II
II X
41
(c > 0). The form [x, y] remains continuous in the new norm
" too:
[x,y]I =I [x+,y+]+ [x-,y-]I =I(x"T)-(x-,y-)I =1(x+-x-,y)1 'lx+-x-11-11Yll"s 11x11-11y11-
(x,YE
)
(6.11)
(here we have used the G-orthogonality of the subspaces ,Y± = P+-.; cf. Theorem 6.4). Since w' is dense in its completion . if- relative to the norm so the form [x, y] can be continuously extended on the whole of 'W_ (we
11.11
keep the same notation for it). We denote by J the closures of the lineals ./t'± (C ') relative to the norm 11 11 " respectively. Moreover, by (6.11),
.+ [1]- relative to the extended form [x, y] and, obviously, 'W+ -relative to the scalar product (x, y ), so that . ie_ =
+ [ O+ ]
- is a Krein space
relative to the form [x, y], or, what is the same thing, it is a J-space, where J = P+ - P- and P± are the orthoprojectors from on to .± respectively. It is clear also that Jis the extension, by continuity relative to the norm II of the operator P+ - P- from ,Y on to the whole of iie- (continuity follows from II
(6.11)).
The question examined in Proposition 6.14 about the imbedding of G-spaces in a Krein space has also another aspect, which reveals, so to speak,
the universality of Krein spaces among all W-spaces. In fact, if w is a W-space, then without loss of generality we may suppose that 11 W11 < 1 (for
otherwise we can renorm N' equivalently, by replacing the original scalar
product ( , ) by a new one: ( , ), = a(-, ), where a > 11 W11; and then for the new Gram operator W, = (l/a)W we obtain 11 W, 11, = (1/a) 11 W11 < 1). Now the space ie_ = Y O+ W is turned by means of the canonical symmetry operator (see Example 3.9)
J=
into a Krein space, and
W
(I - W2)' 2 .
is
(I-
W2)"2
-W
a subspace of it (more precisely,
it is
)-isometrically isomorphic to a certain subspace of it). Since if, in its turn, is (J, J, )-isometrically isomorphic to a certain subspace of an arbitrary i with subspaces X; of sufficiently large dimension, J,-space ,)Y, = .i O+ so . too is (W, J, )-isometrically isomorphic to a certain subspace of the space in its (provided the subspaces ,. Thus, an arbitrary Krein space (W, .
canonical decomposition are of sufficiently large dimension) contains a subspace which is (W, J, )-isometrically isomorphic to a given W-space J(. The theory of semi-definite lineals and subspaces (in particular, the concepts of the intrinsic metric and uniform definiteness, of regular and singular lineals) which has been developed in §§4-6 for Krein spaces also remains valid to a considerable extent for W-spaces and, in particular, for G-spaces.
42
1 The Geometry of Spaces with an Indefinite Metric
Exercises and problems 1
Consider the Krein space .k from Example 3.11 and find a lineal which is dense in this space and indecomposable. Hint: Use the idea of Example 1.33.
2
Prove that the indecomposable lineal (space) .t constructed in Example 1.33 not only is not a Krein space, but (in contrast to the situation in Exercise 1) it cannot in general be imbedded in any Hilbert space ,Y with a scalar product 'majorizing' the form [ . ]: I [x, y] 12 < (x, x)(y. y), i.e., I [x, yl I (11 x 11 11 y 11 (x, y E .t) [VIII]. Hint. Arguing from the contrary, consider vectors x = (i;;); _ with E-k = 1 and = 0 when j ;e -k (k=1,2 ....), and also a vector y = (. . ., 0, ..., 0, 710, nI.... ), where nk -i = (xk, xk) + k (k = 1, 2, ...), and obtain the contradictory inequality qk- i < 11 y 112 for all k = 1, 2, ... (Ginzburg).
3
Let Gy- be the Gram operator of a subspace ' (C ). Then Y is non-degenerate if and only if 11 x 11 v' = 11 Gvx 11 is a norm in 9' ([VI]; see Definition 7.13 below).
4 A sequence
is said to be asymptotically isotropic in a lineal 91 (C.Ye) if
y] -* 0 (n - co) uniformly relative to all y E 2'with II y II = 1. Let G v- be the Gram
5
operator of the subspace Y. The sequence JxJ' is asymptotically isotropic in 9 if and only if lim, - 11 Gvx 11 = 0 (Ginzburg; see [VIII]). A definite subspace Y (C. ,Y) is uniformly definite if and only if an arbitrary asymptotically isotropic sequence in Theorem 7.16 below).
' converges to zero (Ginzburg; see [VIII], cf.
6
The result of Exercise 4 can be substantially generalized: the closure £' of a lineal 9' (C.) is a Krein space if and only if every asymptotically isotropic sequence in converges to zero (I. Iokhidov, see [VIII], Proposition 5.5, and Theorem 7.16 below.) Hint: Use the results of Exercises 3 and 4.
7
Let Y = Y+ [ O+ I Y- [ O+ ] f° be the canonical decomposition of a subspace L' of a Krein space, and let Gv,' be the Gram operators for Y± respectively. In order that
every uniformly positive (uniformly negative) subspace in 2' should be finitedimensional it is necessary and sufficient that Gv-- E 1'm (G, E .y'.) (Azizov).
Hint: In the 'necessary' part use the fact that (iv- - E.,-(k))Gv- - Gv-- when (0< ) 0 in the uniform operator topology, and dim(I, - E, - (X)) < oo when k > 0 (here X Ev-(k) is the spectral function of the operator Gv.-). In the 'sufficient' part use the injectiveness of the orthoprojector P+ from I on to 91+ on positive subspaces (see Lemma 1.27) and the fact that, when x E Il+ (any uniformly positive subspace from f') (Gy--P+x, P+x) >, c11 x112, i.e., Gr- is invertible on P+,. II+.
§7
J-orthogonal complements and projections. Projectional completeness
1
We recall that we have already encountered from time to time J-
orthogonal complements (and even earlier Q-orthogonal complements), starting in §1.5 (see also §1.10, §3.4, §4.5, etc.) But here we begin their systematic study as applied to lineals and, in particular, to subspaces of a Krein space .W with a fixed canonical symmetry operator J in .Y (see §3.5). As before, we indicate J-orthogonality by the symbol [1] and a J-
§7 J-orthogonal complements and projections
43
orthogonal complement by the same symbol raised as an index, while the usual
(Hilbert) orthogonality with respect to the scalar product (x, y) = [Jx, y] is denoted by the symbol 1 and the usual orthogonal complement is indicated by the same symbol used as an index. It is not difficult to prove the Propositions
7.1-7.5, listed below, by direct verification using the relation (3.5) or Propositions 3.5 and 3.7; we leave the reader to do this. ) its J-orthogonal complement [1]) and 7.1 For any subset Af (C its orthogonal complement (.,tll) are subspaces, connected by the formulae .,11[1] = J-11-1,
-111 =
7.2
(7.1a) (7.1b)
For any set /l (C ,W) Ill [1] =,tl [1] = (Lin lf) [1] .
7.3
For any
If (C) (7.2)
(J-11)[1] = 7.4
For any set -1l (C) .4,1 1-11
[1]
= C Lin ill
(cf. (1.11)), and, in particular, for any lineal Y (C W) Y[1] [1]
7.5
_ g.
(7.3)
For a subspace q (=.Ii) it is always true that Y[1] [1] = 2, and
therefore the isotropic subspaces 9° of 2 and T 1-L1 coincide:
2° _ Ynr[1] _ -r[1] n-v[1] [1].
In this paragraph we restrict our attention to the case mentioned in 7.5 above, where ' = 2' is a subspace of a Krein space. Although it is always true 2
that 2'n 21 = [B1 and 911 + 2' = Y relative to the Hilbert metric, yet the analogous equalities with respect to the indefinite metric are, generally, speaking, not true if only because 2' may be a degenerate subspace, and then 9n 2'[1] _ 2'° (0). In this case the algebraic sum 99 + Y[1], i.e., the lineal ' + Y[1] _ (x, y I X E 2', y E 2'[1] 1 is no longer a direct sum.
7.6 For any subspace 9 the relation (J2'°)1 = (2'+ 2'[1]) holds. Using (1.12) and Proposition 7.5 we have (2' + 2[1J) [1] _ 2'[1] n 2'[1] [1] = 2'°, whence, by (7.16), (2' + 2'[1] )1 = J2'°
and (J2'°)[1] = (2+ 2'[1])11 = (2'+ 2n ).
U
1 The Geometry of Spaces with an Indefinite Metric
44
The fact proved in Proposition 7.6 implies that the whole space . expressed in the form of an orthogonal sum: .,Y = (2' + Y11]) O+ J2'°.
can be (7.4)
This enables us to replace the relation 2 @ 2i1 = .A', which is true only in the Hilbert metric by the following more precise form of it: Lemma 7.7:
In order that Y+.x[11 =.W'
(7.5)
it is necessary and sufficient that the subspace .9' be non-degenerate.
The proof follows directly from formula (7.4) if we take into account that, since the operator Jis unitary (see (3.4)), the relations J2'° = [0} and 2j0 = (0} are equivalent.
Returning to the question discussed in Paragraph 2, but from a more general viewpoint, we introduce the following 3
Definition 7.8: A lineal .' in a Krein space is said to be projectively complete if Y + g11] _ W.
(7.6)
In seeking conditions for projective completeness we discover that, as in the
analogous question in the case of a definite (Hilbert) metric, the following proposition holds: 7.9
For a linear 2' (C W) to be projectively complete it must be closed: 2' = 2'.
In particular, the subspaces 2' and Yf 1 ] can be projectively complete only simultaneously. El Suppose (7.6) holds, but there is a vector yo E 21191. We express the vector yo
in accordance with (7.6) in the form yo = xo + Zo (xo E 2', zo E 2111). Thus, xo, yo E2', (yo - xo = )zo E 2 n 2r11] = y[1] [1] n 211] , i.e., zo [1] SY + .9'[l] =,w
and zo = 0; hence yo = xo E . contrary to the choice of yo. The second assertion in 7.9 follows from the equality Y111 11] = go (see Proposition 7.5). However, although in the `definite' situation the condition 2' = 2' is not only
necessary but also sufficient for the existence of the decomposition .7Y = 2' U x'0, the indefinite metric, as already mentioned at the beginning of paragraph 7.2, here shows its specific character. In particular, we have 7.10 In order that a subspace 9' (C W) should be projectively complete it must be non-degenerate: 2' n 241] = [0}. This is a simple consequence of Lemma 7.7, just as (7.5) follows afortiori from (7.6).
§7 J-orthogonal complements and projections
45
But even both the conditions 9' = 2' and y, fl M'I1I = (B) together are not
sufficient for 9' to be projectively complete. We can quickly see this by approaching the problem from a rather different standpoint. 4 A vector x is called the J-orthogonal projection of a vector y E .,Y on to a subspace 2' (C.R') if 1) X E M'
2) y - x [1] Y.
(7.7)
For every (so to say, `individual') vector yo a criterion for the existence of its J-orthogonal projection on to a given subspace 2' (C ) is obtained simply in terms of the Gram operator G, of the subspace 2' (see Definition 6.3). Lemma 7.11: Let P,, be the (Hilbert) orthogonal projector on to a subspace £' (;4 (B)) of a J-space W and let G,- = P,(J 1 2') be the Gram operator of the subspace Y. Then in order that a vector yo (E W) should have a J-orthogonal projection xo on to 2' it is necessary and sufficient that PvJyo E dRG,.
The existence of the required J-orthogonal projection xo(E rL') of the vector yo is equivalent, by (7.7) to [yo - xo, x] = 0 a [yo, x] = [xo, x] a (Jyo, P,x) = (G,xo, x)
holding for all xE 2, i.e., (P,Jyo, x) = (Gvxo, x), whence P,Jyo = Grxo E 3?G,. Since the whole argument is reversible, Lemma 7.11 is proved.
Lemma 7.11 does not answer the question whether the J-orthogonal projection is unique when its projection exists.
In order that, even if only for a single vector yo (E.) with a J-orthogonal projection xo on to a subspace 2' (;4 10) ), this J-orthogonal projection should be unique, it is necessary that 91 be non-degenerate. This same condition is sufficient to ensure that any vector y (E,') shall have not more than one J-orthogonal projection on to 2. Lemma 7.12:
The existence of an isotropic vector zo (,,60) in 9' would bring it about that, for the vector yo (E W) which has the J-orthogonal projection xo on ', the vector xo + zo (* xo) would also be a J-orthogonal projection on to d', because xo + zo E 2' and yo - (xo + zo) [1] Y. Conversely, if for some vector y (E.*') there are two orthogonal projections x, and x2 on to 2' with x, ;d x2, then the vector (0;4 )zo = x, - x2(ESP) has the
property: zo = (y - x,) - (y - x2) [1] 2', i.e., it is an isotropic vector for 9.
5 We now return to the search started in para. 7.3 for criteria for the projective completeness of a subspace 99 (C.), and for this purpose we introduce some further definitions.
1 The Geometry of Spaces with an Indefinite Metric
46
Definition 7.13:
In a non-degenerate subspace 2' having the Gram operator
G r we introduce the norm 11 x lI y' = II Gyx II (x E 2').
Definition 7.14: A non-generate subspace 2' is said to be regular if the norms 11.11 r and II II are equivalent on it. -
7.15 A subspace 91 with a Gram operator Gr is regular if and only if 0Ep(Gy). E The estimate II G yx II = I I x I I y' > c 11 x 11 for all x E 2 (c > 0) is equivalent
to the continuous invertiblity of the bounded selfadjoint operator Gy-. Now we can establish a fundamental theorem, containing a set of criteria for projective completeness. Theorem 7.16: For a subspace (101;4) 2' (C W) with a Gram operator Gy the following four assertions are equivalent:
a) 2' is projectively complete; b) 2' is regular, i.e., 0 E p(Gv) (see 7.15); c) 2C' is a Krein space;
d) any vector y E ,' has at least one J-orthogonal projection on to 21. a) b). Since 2' is projectively complete, it follows from 7.10 that (7.6) represents a decomposition into a direct sum: W = 9? [+] To this decomposition correspond the bounded projectors Q and (I - Q) respectively. Therefore for any (0;4 ) x E 2' and y = (1 11 x 1I )Jx we have El
Ilxll = [x,y] = [x, Qy] =(Gyx,y) < II G1x1111 Q11
= II QII II xIlrs II Grl1 II QII lixII, i.e., .' is regular. b) c). This implication was established earlier in Corollary 6.12. c)
d). By Theorem 5.7 2' = 9?+ [+] 2-, where 2`- are uniformly definite. We
show that any vector y E .,Y has a J-orthogonal projection on to 2+ and 2-. Consider, for example, a linear functional (py (x) = [x, y] (x E 2+ ). Since I oy(x) I = I (x, Jy) 15 11 x 1111 y 11, so ivy is continuous relative to the norm 11 11, and therefore also relative to the intrinsic norm I I y- which is equivalent to it (see
(5.2) and (5.3)). Since 2+ is complete relative to I y- (see 5.6), there is, by Riesz's theorem, an x, E 2+ such that (x) _ [x,- x,] (x E 2+ ); hence I
[x, y] = [x, x,] for all x E 2+ and y - x [I] Y'. Thus, x, is the J-orthogonal projection of y on to 2+, and similarly we find an x2, the J-orthogonal projection of y on to 2-. But then xo = x, + x2 is the J-orthogonal projection of
y on to 2'. d) - a). Since any y (E.) has a J-orthogonal projection x on to 2, so y = x + z (x E .', z E 2I') and W = 2' [+] 2'-, i.e., 21 is projectively complete. Corollary 7.17: If a subspace 2' is definite, then its projective completeness is equivalent to uniform definiteness.
§7 J-orthogonal complements and projections
47
This follows, for example, from 7.10, 6.11, and the equivalence a) = b) in Theorem 7.16. Corollary7.18: Every finite-dimensional non-degenerate subspace 9? is projectively complete.
This follows from assertion b) in Theorem 7.16 and the equivalence of all norms in a finite-dimensional space.
To conclude this section we again return to the question about different canonical decompositions of a Krein space .' and the Hilbert topologies 6
determined by them (cf. Remark 2.4 above). We consider, as well as a particular fixed canonical decomposition
.) = J+ [--].'-
(7.8)
with the canonical projectors P+-, the canonical symmetry operator J = P+ - P-, and the scalar product (x, y) = [ Jx, y] and norm II X11 = (x, x) v2 (x, y E .-W) generated by them (see §3), some other canonical decomposition
.0 =.Wi
(7.9)
i.e., we carry out, as it were, `a rotation of the co-ordinate axes'. It is clear (7.9) also generates corresponding canonical projectors Pi : P;~ "= . i , P1 + Pi = I, a new canonical symmetry operator J, = P1 - Pi , and also a new scalar product (x, y), = [ J1 x, y] and norm II xII, = (x, x)1 12 (x, y E *). Theorem 7.19: The norms II ' 11 and ' 11, generated by different canonical decompositions (7.8) and (7.9) of a Krien space ,W are equivalent. II
We start from the fact that one of the canonical decompositions, let us say (7.8), generates in Y (in accordance with §2.2) the structure of a Hilbert space with the norm 11 x 11 = (x, x)1/2 (x E .°). The presence of the other canonical decomposition (7.9), where .1i are definite lineals and Wi [1] _Yi , shows that they are projectively complete, and closed in the norm II ' (see 7.9), that the projectors Pi are closed in this norm, and finally that XP are uniformly definite II
in this norm (see Corollary 7.17). If we now take into account that the norm II
II,
is the intrinsic norm on ,i (cf. (5.1)), then it is clear that on.Yi the norms II ' and II ' 111 are equivalent. Therefore i , and with them also the whole of . (see 2.3), are complete relative to the norm 11 ' 111. We remark further that for x1 E .Wi the norms II x1 111 (as intrinsic norms) II
are simply subordinate to the original (external) norms: II x; II 1 S II xi II Therefore, for any x E , x = xi + xi (xi E . 'I) we have 11X111=IIx; +x1-II<, IIx1 III+11x1 III
1 The Geometry of Spaces with an Indefinite Metric
48
where
O= 11 P1 II+IIP1II>0. Hence by virtue of the completeness relative to both the norms I it follows that these norms are equivalent.
'
II and II '
Exercises and problems A semi-definite lineal 2' is uniformly definite if and only if 9 is projectively complete ([VI)2 Let 2' (C.w°) contain a maximal uniformly positive (uniformly negative) lineal (see Exercises 3 and 4 to §5). Then 2' is projectively complete and 2'' is uniformly negative (uniformly positive) ([V] ). Hint: Use Corollary 7.17, Exercise 4 to §5, Proposition 5.3, and also 7.9 and 7.4. 1
3
A lineal 2' is a maximal uniformly definite lineal if and only if there is a cononical decomposition w' =,01' [+] . ,- such that 2' = JYt (or 2' = ,Wj) ([VIII]). Hint: Use the preceding Exercise and the result of Exercise 3 to §5.
4 A lineal 2' is projectively complete if and only if a) 2' is closed b) 2' is non-degenerate c) for any decomposition of it 2' = 2P' [+] 91- (2+ C U [9l) there is a canonical decomposition . Y = 1 [+] .; °j such that 2+ C .0; (m). Hint: In the `necessary' part use 7.9, 7.10, Theorem 7.16, the maximality principle for uniformly definite lineals (see §5, Exercise 3), and the result of Exercise 3 above. 5
In Theorem 7.16 deduce the equivalence d) a b) from Lemma 7.1.
6
Without using Theorem 7.16 prove that for subspaces 2' which are definite, regularity is equivalent to uniform definiteness. Hint: Compare the norms 11 x 11 r-=11 G rx 11 and 1 x 1 v =11 G 12x 11 (x E r, Y' is positive). Let .Yt° be a Krein space, and 2' be a subspace of it which admits the decomposition Y'= o [+] 99', where 2' is positively complete, and Y'o is the isotropic part of 2'. Then in any decomposition 2lll = 2'o [+] (-W[ -L])' into a direct sum of subspaces the second term will be a projectively complete subspace.
7
Hint: Use the result of Exercise 10 to §2, taking into account that (9911))'
is
J-isometrically isomorphic to the subspace MI-L1/2'o. 8
Prove that the sum 2, [+].r2 of projectively complete subspaces 2, and 2'2
is
projectively complete.
§8
The method of angular operators
I Let 2'C -' in a Krein space W _ ,Y' [ O+ ] . If P+ are the canonical projectors, then by Theorem 4.1 the mapping P' 12': 2' P'1' (.W') is a linear homeomorphism, and II (P' I Y)-' II J. We now consider the
bounded linear operator
K= P- (P+
K: P' '-.
-.
(8.1)
§8 The method of angular operators
49
Definition 8.1: The operator .yf defined by the relation (8.1) is called the angular operator for 2' with respect to y" . The meaning of this nomenclature is explained a little later (see Exercises 1 and 2 to §8).
For an arbitrary vector x(E2') we have x = x+ + x- (x± E .Y±), x + = P+xE P+5, Kx+ = P-(P+ j2')-'x+ = P-x= x-. Thus every vector x (E2) has the form x = x+ + Kx+ (x+ E P+2'), and since 2' is non-negative, for all x+ E P+2' we have (see Proposition 2.5) II x+ x Kx+ 11, i.e., I < 1 (K is a compression). KI Conversely, given an arbitrary 2+ (C.Y+) and an arbitrary compression K: 2+ (11 K j < 1), we consider the set of vectors .' = (x+ + Kx+ }X- E,,,-
(8.2)
Since K is a linear operator, 2' is a lineal and 2' C P+ because, for any x = x+ + Kx+ (E2') we have x- = Kx+, hence 11x 11= I Kx+ 11 < 11 x+ 11. Moreover it is clear that P+2' = 2+ and K is an angular operator for 2' with respect
to M+. Summing up, we have obtained a complete description of all non-negative lineals of a Krein space, and, in fact, we have proved The formula (8.2) in which 2+ is an arbitrary lineal from+, and K: 2+ . - is an arbitrary compression (II K 11 < 1), gives the general form of all 2' C + of the Krein space Y = W + [ O+ ] ,Y-, and 2+ = P+9 and Theorem 8.2:
K is the angular operator for 2' with respect to .W+.
All non-positive lineals 2' (C.) are described similarly. Definition 8.1': If 2' is a non-positive lineal in the Krein space _Y = ,Y+ [ ] . , then the operator
Q= P+(P- I `')-',
Q: P-2'-+ ,Y+
(8.3)
is called the angular operator for 9.9 with respect to .0-. The reader will be able himself to establish the following analogue of Theorem 8.2:
Theorem 8.2':
The formula
+X XE,' 2= (QxJ E , in which 2- is an arbitrary lineal form .W-, and Q: 2- - .W+ (11 Q 11 < 1) is an arbitrary compression, gives the general form for all .' C .?- in the Krein space .0 = .,Y+ [C+ ] .-W-, and 2- = P-'2, and Q is the angular operator for 2' with respect to 2
From now on (Sections .2 and .3) we restrict ourselves to the case of lineals
9' C -+ and their angular operators K, since for 2' C .0- the whole theory is
1 The Geometry of Spaces with an Indefinite Metric
50
entirely analogous (the reader can himself formulate the corresponding Proposition 8.3', Lemma 8.4', etc.). In spite of its simple construction the angular operator K for 9 (the brevity we shall often omit the words `with respect to .,Y') contains quite a lot of information about the properties of the lineal itself. 8.3
If K is the angular operator for 2'C
.-+,
then 2'C .-++ U 10) if and
only if II Kx+ II < II x+ II for all (6 ?6 ) x+ E P+2, and .' C ,i°° if and only if K is an isometric operation: II Kx+ II = II x+ II (x+ E P+-W) Since for x E 91 we have x = x+ + x- = x+ + Kx+, all the assertions follow immediately from the formula (see (2.7))
[x,x]=IIx+III-IIx III= IIx+II2-IIKx+III.
We note that for 2 C e° the description can be obtained by means of the
analogue 8.3' of Proposition 8.3 in terms of the isometric operator Q: P-91 - ,Y+ which is the angular operator of 2' with respect to ,Y- (see (8.3)). Here, obviously, Q = K-'. A rather more subtle fact than Proposition 8.3 is revealed by Lemma 8.4:
2' (C 1 +) with the angular operator K is uniformly positive if
and only i f 1 1K1 1 < 1.
Let 2' be uniformly positive, i.e., for all x E 9'
IIx+III- IIx- III= [x, x]>a'IIxlll=al(IIx+III+IIx-III)
(a>0), (8.4)
and by the meaning of Definition 5.2 a can always be chosen so that 0 < a < 1, and this we do. The relation (8.4) is equivalent to the inequality
(1-a2)Ilx+III> (1+a2)IIx
112,
(8.5)
whence 2
KII
IIKx+III=IIx III
2<1. 1
Conversely, let II K II = p < 1. Put
a
1-p2 T+
(0
p
P2 = (1 - a2)/(l + a1), Then x = x+ + x- = x+ + Kx+ we have
and
for
all
x+ E P+2'
and
III=IIKx+112 2llx+III=1+aIIIx+III,
IIx
or (1 - a2) II x+ > (1 + a2) II X- III, which, as we have seen (cf. (8.5)), is equivalent to (8.4). III
§8 The method of angular operators
51
Lemma 8.4 together with Proposition 8.3 shows an effective method of constructing both definite and uniformly definite lineals.
The concept of an angular operator proves to be particularly useful for work with maximal semi-definite subspaces. For brevity of exposition we introduce the following two sets of subspaces: 3
(,It' (.A-) = )-//+-the set of all maximal non-negative subspaces of a Krein
space,
(8.6)
(.,11- (Y) = ).,tf--the set of all maximal non-positive subspaces of a Krein space W. (8.7) In order that 2' (C P+) be contained in it is necessary and sufficient that its angular operator K be defined everywhere in 0+ (K: 4e-). This follows from Theorem 4.5 and 8.2. Matters are naturally more complicated with maximal positive lineals 2' 8.5
because as we have already seen (see Example 4.12) such an 2' may be non-closed and in this case .' ¢ W+. 8.6 In order that 2' (C.?++ U {O)) with the angular operator K_ should be a maximal positive lineal it is necessary, and if 2' is closed (2' = 2') it is also sufficient, that the domain of definition of the operator be dense in . +:
'K=.+.
This is simply a re-formulation of Theorem 4.7.
Propositions 8.5 and 8.6 outline a way of constructing extensions of semi-definite and definite lineals to maximal lineals of the corresponding classes, and, moreover, a way of describing all such extensions.
8.7 Let 2' C (C.W+). Then:
P+, and let its angular operator K be defined on the lineal VK
a) non-negative extensions 2' (3 2') are described in accordance with Theorem 8.2 by all possible contractions k (3 K), i.e., by extensions of the operator K with II K II < 1 for which t.)i C W+ and K: LAg-' .q-; b) in particular, all . (3 2') from ,,ff+ are described by such contractions k (c K) for which CAk = ,Y+
This follows from Theorem 8.2 and Proposition 8.5. Remark 8.8: The procedure for constructing 2' (E,11+) from assertion b) of Proposition 8.7 begins with the closure of 91, i.e., consideration of the non-negative subspace SE, with the angular operator k (the closure of the operator K). If 1K = '+, then it is clear that k E .,tf+ and there are no other `/' E tt+ such that 2' 3 Y. But if LAK ? 4Y+, then one of the extensions 2'
(3 2') of the class -tt+ can be constructed (without resorting to Zorn's
1 The Geometry of Spaces with an Indefinite Metric
52
lemma-see Remark 1.20) for example by means of the trivial extension k of the contraction K on to the whole of +, namely by defining the linear operator K by the formulae x+ _ (Kx+, x+ E 99, B,
X+ E (/+ (=.,Y+ (D cK)
(8.8)
11
But as regards assertion a) of Proposition 8.7, here intermediate extensions 2' of the lineal 2' between 2' and 2' are possible (2' C 91 C SP).
Remark 8.9: The procedure described in formula (8.8) of Remark 8.9 is applicable also to the construction of maximal positive extensions 2' of a positive 2' with the angular operator K, but only in the case when its closure 2' is also positive.
4
Turning to the consideration of neutral extensions of neutral lineals we need
(in accordance with what was said after the proof of Proposition 8.2) the Theorem 8.2' and also the analogue 8.7' (which we have not formulated) of Proposition 8.7 for non-positive lineals. Theorem 8.10: Let W = + (e ] W- be a Krein space, let 2' C P °, and let K and Q be its angular operators with respect to -Y+ and .YY- respectively. Then:
a) all neutral extensions . ' of the lineal 2' are described in accordance with
Theorems 8.2 and 8.2' by all the angular operators from the union (K) U (Q) of the set (K) of all isometric extensions k (3 K) of the isometric
operator K (1gC
W+, K: W - W-) and the set (Q) of all isometric
extensions Q (3 Q) of the isometric operator Q (#1 C AY-, Q: CAQ - .Ye+ );
b) in particular, all maximal neutral extensions 2' (3 2') in the case dim Ae+ < dim .-W- are described by all the maximal isometric operators k (Vg= W+) described in assertion a) of the class (K), and in the case dim +
> dim W- by all the maximal isometric operators Q mQ = .W- )
described in assertion a) of the class (Q); c) in the case dim W+ = dim M- = oo the description of all maximal neutral extensions 2'(3.') is obtained by considering all the maximal isometric extensions of the operators again from the union (K) U (Q) in assertion a); but in the case dim + = dim M- < oo it suffices to use the maximal isometric operators only from (K) or only from (Q);
d) hyper-maximal neutral extensions 2' (3 2'), which exist only when dim W+ = dim ,Y- are described by the set of all unitary extensions k (3 K) (or, what is equivalent, Q (3 Q)), i.e., by isometric extensions such that y- (Q.W = .YY+ ). Assertion a) follows from 8.3 and 8.3'.
§8 The method of angular operators
53
b) If dim W+ < dim W-, then every maximal neutral subspace (D-T) belongs to the class -ft'. In the opposite case in accordance with Corollary 4.14
we would have 9 E Jr, P-2 _ 4e-, and on the basis of the analogue 8.7' of Proposition 8.7 b) the angular operator Q of the subspace 2' with respect to Wwould be defined everywhere in M- and would be isometric, and a. MY+; but this is impossible when dim + < dim,*-. It remains only to refer to 8.7 b). In the case dim W+ > dim Y- the argument is similar. c) In view of the isometricity, when dim M+ = dim .-, of the Hilbert spaces
,W+ and W-, in the case when they are finite-dimensional all the maximal isometric (unitary) extensions K: M+ on to W- of the operator K and the maximal isometric (unitary) extensions Q: Y- -on to M+ of the inverse operator Q can be obtained from one another by the formulae Q = K-' (or K= Q-'). But if the dimension dim W+ = dim W- is infinite, then maximal isometric extensions K:
+ - W- for which K + # W- are possible, and
similarly Q:.W -+M+ with Q."- ;4 ,2+ are possible. Therefore a complete description of all maximal neutral extensions 9 (J SP) is achieved only by a sorting-out of all the maximal isometric operators from the union (K] U (Q]. Of course, with regard to those of them which are unitary, we can limit ourselves in this sorting-out, as in the 'finite-dimensional' case, to the unitary operators from (K} alone (or those from (Q] which are inverse to them). d) The truth of the last assertion of the theorem follows from 8.7 (or from 8.7') and Corollary 4.14.
In this section we consider only subspaces 2' from the A+ class and the /!/class. Therefore their angular operators K (resp. Q) are defined everywhere in 5
.)+ (resp. Y-). We now need the well-known concept of the operator T* adjoint to an operator T. lei - 2 acting from one Hilbert space (,',) to another (,'2) (see, for example, [VII]). Later (see Chapter 2, §1.1) this concept will be significantly generalized, but in the simplest particular case, which is what we now need, when the operator T is bounded and defined everywhere in .0,, the adjoint adjuster T*: .1Y1 is naturally defined on all y2 E W2 by the relation -IY2 (x1, T*Y2)I = (Txi,Y2)2
(xi E',),
(8.9)
where
and ( , )2 are the scalar products in A", and .2 respectively. It is clear from (8.9) that T* is also bounded and that it has the usual properties: (S+ T)* = S*+ T*,
(ST)* = T*S*,
T**= T.
Theorem 8.11: Let 2' E -&+ and let K be the angular operator of 2' with respect to .Yt'+. Then 2'- E mil(- and the angular operator Q of the subspace 2'1l] with
respect to W- is K*: Q = K*.
(8.10)
1 The Geometry of Spaces with an Indefinite Metric
54
El Let y E Y,f1] . Then, for all x E L, we have
0= [x,Y]= [x++Kx+,Y]=(x+,Y+)-(Kx+,Y
)=(x+,Y+)-(x+,KY
),
(8.11)
i.e., for all x+ E.J+ (x+, Y+) = (x+, K Y )
(8.12)
(here the roles of the spaces . , and A e2 in the definition (8.9) are played by ye+ and W- respectively). But since (cf. (8.9)) K*y- E 3e+, it follows from (8.12) that
Y+ = K*y- and so y = K*y- + y-. Conversely, for all y- E W- the vector y = K *y- + y- (y+ = K *y-) satisfies the relation (8.12), from which, taking the steps in the reverse order, the chain of equalities (8.11) is established. Thus y E 27111. So
Y"1 = [*y- + Y- ]y- (.H- and I K* II = II K II < 1 so that X111 E /#- and its angular operator Q = K*.
Theorem 8.11' (the dual of Theorem 8.11) the reader will be able to formulate for himself without difficulty.
Remark 8.12: For any subspace 9? (=2), as is well-known (see 7.5), we + always have Y[11 [tl = Y. In the particular case when 2 E to this fact the following relations, deriving from Theorems 8.11 and 8.11', for the corresponding angular operators are equivalent: K**= K (Q**= Q). Corollary 8.13 (8.13'): If 2 E /u+ (Jr) and is positive (negative), then the subspace Y[11 E elf- (.,ff+) is negative (positive).
Suppose, for example, that 2 E . lf+ and is positive. Then, by Theorem 8.11, x[11 E tf-. But also x[11 is non-degenerate, for otherwise any isotropic
vector zo (;r-1 B) in it would be J-orthogonal to 2 and then 9 = Lin [ 2', zo 1 (;4 2) would be non-negative and wider than 2' (zo ¢ 2, because 2' is positive), which contradicts Corollary 4.8.
Corollary 8.14 (8.14'):
If 2' E ff+ (Jr) and is uniformly definite, then
Wi[t] (E_ff ) (x[1] E off+) is also uniformly definite. Moreover 99 [+] 2111 =,)(e.
7 This follows from the fact that II K* II = II K II < 1 (II Q * II = II Q II < 1) for the angular operators K and K* (resp. Q and Q*) of the subspaces 2 and 99[11 respectively. The last assertion follows from Corollary 7.17
For the pair of subspaces .M'± E ,.ff+ it is possible to obtain a curious characterization (widely applicable later) of the mutual disposition of these 6
§8 The method of angular operators
55
subspaces in terms of the spectra produced by their angular operators K and Q respectively. As a preliminary we recall that the operators QK and KQ act in the Hilbert spaces .-W+ and - respectively. Theorem 8.15:
The following implications hold:
a) T+ n Y_ 5e (B] a 1 E up (KQ) U up (QK); b) T+ [+]2_ =. a ap(KQ)Uop(QK);
c) -'+[+]Y-=-ye ap(KQ)np(QK) a) For any vector xo E '+ n Y_ we have two representations:
O
xo = xo + Kxo = Qxo + xo
(xo E W± , xo + xo = xo ),
from which it follows that
Kxo =xo.
Qxo =xo, Hence
(8.13)
KQxo = xo
Qkxo = xo .
and
0 also, and so
If xo P6 0 it follows from (8.13) that xo
1 E ap(KQ) n up_(QK)(Cup(KQ)Uup(QK)).
(8.14)
Conversely, if, for example, QKxo = xo (;40), then (0 ;4 ) Xo + Kxo = QKxo + Kxo E Y+ n _r_
b) The inequality 2'+ + 2_ # W is equivalent to the existence of a vector (0 ;6 )yo L 2+ + Y_, or, what is the same thing, xo = Jyo [1] Y+ + Y_. Since xo [1] Y+, so [xo, xo] < 0; and it follows from xo [1] 2'_ that [xo, xo] > 0, i.e., xo is a neutral vector. Since 2'± are maximally semi-definite, it follows that
xo E9,, n2'_, and the whole of this chain of reasoning is reversible. Thus
2'+ + 2'_ # if o 2+ n2'_
[e],
and
it
remains
to
use
assertion
a)
(in particular, the `principal part' of the relation (8.14)). c) Let 2+ [+] 2'_ = W. Then for every x+ E W+ there are xo E -W± such that
x+=xo + Kxo +Qxo+xo.It is clear that x+=xo +Qxobut Kxo+xo =B, and therefore for any x+ E W+ the equation xo - QKxo = x+ is soluble (with respect to xo E
+ ). Therefore the number X = 1 cannot enter into either
the continuous spectrum nor the residual spectrum of the operator 1 E ap(QK)Up(QK). But 2' +n2'_ = [9}, so it follows from a) 1 E p(QK). Similarly we can prove that 1 E p(KQ), i.e., I E p(QK)np(KQ) C p(QK)Up(KQ). Conversely, suppose, for example, that 1 E p(QK). Since an arbitrary QK:
that
vector from 2+ + 2- can be written in the form x+ + Kx+ + x- + Qx-, where x± E ±, it is sufficient to verify that, for any y' E W±, the system x+ + Qx- = y+, x- + Kx+ = y" is soluble. From the second equation x- = y- - Kx+, and so we can -rewrite the first equation in the form
1 The Geometry of Spaces with an Indefinite Metric
56
x+ - QKx+ = Y+ - QY-, and therefore x+ = (I+ - QK)- (y+ - QY ) Hence QK)-' (y+ - Qy- ).
x- = Y - K(I+ -
Corollary 8.16: If in the pair Y± (C J±) at least one of the subspaces Y± is uniformly definite, then T+ [+] 2'_ = W.
This assertion follows from Lemma 8.4 and assertion b) in Theorem 8.15.
7 We now return to the investigation started in §7.6 of the question of a `rotation of the co-ordinate axes' in a Krein space W, i.e., of a transition from one of its canonical decompositions
.) =
(8.15)
+ [+] .Y
to another decomposition . = Yei [+] Ml-. We shall regard the decomposition (8.15) with the canonical projectors P± (- 0) and J = P+ - P- as fixed and as defining the scalar product (x, y) = [ Jx, y] ((x, y) E Jr'), )and we pose the problem of calculating the new canonical projectors Pi : W - Yi . We recall that Mi are uniformly definite and for their angular operators K and K* respectively (see Corollary 8.14) we have II K II = II K* II < 1.
Theorem 8.17: The new canonical projectors P± are represented in matrix form relative to the decomposition (8.15) as K*K)_K*
II
II
- K*(I- - KK*)-'K K*(1- - KK*) ' - (I- - KK*)-'K (1- - KK*)-' II'
P' Pl _
-
- (I+ -
(I+ - K*K)-' K(I+ - K*K)-'
- K(I* - K*K)- IK* I I'
(8.16)
where I+- are the identity operators in 4 ± respectively.
The existence of the operators represented by the right-hand sides of formula (8.16) is obvious. The fact that they are idempotent, i.e., (P; )2 = Pi , is verified immediately by squaring the corresponding matrices. Further, the construction of these matrices is such that for any vector-column x
x+
x
E.W
we have
P1 x= P1
x+
x
H
(I+ - K*K) -'x+ - (I+ - K*K) -'K*xK(I* - K*K) -'x+ - K(I+ - K*K) -'K+x-IIKy+11=YE.-Y,+,
where y+ = (I+ - K*K)-'x+ - (I+ - K*K)-'K*x- E ,P+.
§8 The method of angular operators
57
In exactly the same way
(K*z)
Pix=
=zE_Yi
where z- = - (I- - KK*)Kx+ (I- - KK*)-'x- E W-. It remains to verify that y and z are the J-orthogonal projection of the vector x on to .w, and 01 respectively. We have
x- y= _
,
I
x-y X+
K+ +
K*K)-'x+ - (I+ - K*K)-'K*zx+ - (I+ x- - K(I+ - K*K)-'x+ + K(I+ - K*K)-'K*x-
,
so that for any vector u+ I
ku+
Eli
,
carrying out straightforward calculations, we obtain
[x- y, y] = (x+ - y+, u+) - (x- - Ky+, Ku+) =(x+-(I+-K*K)-'x++(I+-K*K)-'K*x-,u+)
-(x--K(I+-K*K)-'x++K(I+-K*K)-'K*x ,Ku+)
([I+ = - (I+ - (-K*K(1+
K*K)-')x+
+ (I+ -
K*K)-'K x-, u+)
- K*K)-'x+ + [I+ + K*K(I+ - K*K)-')K*x-, u+) =([1+-(I+-K*K)-'+K*K(I+-K*K)-')x+
+ [ (I+ - K*K)
I+ - K*K(I+ -
K*K)-')K*x-, u+)
=(0,u+)=0, i.e., x - y [L] .i , and in exactly the same way one proves that x - z [L] .i .
8 We shall now indicate some useful generalizations of theorems using angular operators, and at the same time some generalizations of the concept `angular operator' itself. Suppose, for example, that in a non-degenerate subspace 9 of a Krein space JY the Gram operator G generates the canonical decomposition (see Theorem 6.4)
Y_ Y+ [q)]Y-,
(8.17)
and that the Gram operator G, itself relating to (8.17) is represented by the matrix
110I
-A I'
(8.18)
1 The Geometry of Spaces with an Indefinite Metric
58
where I (= I,--) is the identity operator in
Y+,
and A (> 0) is a bounded
positive operator in the Hilbert space 99-, and it is possible that 0 E oc(A), so
that Y- is a negative, but not, generally speaking, a uniformly negative subspace (cf. Theorem 6.12).
If the subspace - satisfies the conditions (8.17) and (8.18), then any non-positive lineal 2, (C Y) can be expressed in the form 2'1 = (Fix- + x- )X- E P.,-Y- where P,,-- is the orthoprojector on to 2 and F, is a linear operator, F1: Pv--?, Y+ The subspace Y, is a maximal non-positive subspace in 21 if and only if 8.18
. PY'-Y, = We note that, in accordance El x = x+ + x- E 22(x± E 21±) we have
with
(8.17),
(x+, x+) - (A 112X-
[x, x] = (G,'x, x) = (x+, x+) - (Ax-, x-) =
for
(8.18)
,
any
A,i2x ),
and therefore, in particular, for x E 2t, where [x, x] 5 0, II x+ II = (x+,
x+)
S (A vex-, Av2x-) = A vex- II2 II
(x- E P,'-2'1). (8.19)
The operator Q, which relates to any vector A "2X (E A "2P-211) the vector
x+ E Y+ is correctly defined as a linear operator, for in accordance with Lemma 1.27 the vector x+ is uniquely regenerated by x- (E P,--211), and it follows from (8.19) that when = 0 we have x+ = 0. Thus, A,i2x-
x+ = Q,A112x- (x- E Pr-211), and further, again by virtue of 18.19), <, 1, so that F1 = Q1A 1/2 is indeed the required operator. Further, when Py--2', = 2- the maximal non-positivity of 11 in 2- follows from Corollary 1.28. Conversely, let 21 be the maximal non-positive space (in 21). Then, by what has been proved, 21 = (x- + Q, A 12x- )X- E P, v,,. But the contraction Q1 defined on A' 2 P,'- '1 (C 2-) admits extension into a II Q1 II
contraction defined on A 122- which maps A 1/22- into 2+; this would imply a non-positive extension of 21, in Y. But 2, is maximal, and therefore Q, is defined on A 1/2w-. Hence it follows that indeed P,- = M. For, let us now consider ie1 = (x + Q1A,/2X )X-(
'-
(CY);
by virtue of (8.19) this is again non-positive and it contains 9'1, i.e., 91 = Y,
and P,-211 = P,,-21 = 2-. It is clear that a proposition analogous to 8.18 can be established for the case when G,,=
A
0
0
-I
,
A > 0,
and 2, (C 11) is non-negative.
We note further that it is possible by means of angular operators to describe not only semi-definite lineals but a far more general type of lineals of a Krein space ." = ,Y+ [-+].,Y-. With this purpose let us consider, for example,
§8 The method of angular operators
59
the class d+ of all lineals Y from .W' which are mapped injectively into M+ by
It is not difficult to figure out that ,+ consists precisely of a projector those 2' for which 2' fl .- = (0) (compare with the proof of Lemma 1.26). It is clear that all such 2' admit the following description: P+.
2= (X++KX+)x-EP-/', where K = P- (P+ I 2')-', but on this occasion the `angular operator' K is no
longer obliged, generally speaking, to be a contraction. In fact,
it
is a
contraction if and only if 2' is non-negative (it is obvious that all non-negative are contained in d+ ). The class d- of all 2' for which 2' fl + = (0) is defined analogously. They are described by the formula
2'= (QX + X)z EP- I,
where Q = P+ (P- 91)-'. The class d- contains, in particular, all nonpositive Y.
In conclusion we consider `in the large' the sets .Yl ± of all the angular operators K(Q) of subspaces of the class /ll± respectively. From Theorems 8.2, 8.2' and Proposition 8.5 it follows that: 9
8.19 (
+,
+
(
(
-)=) +_ (KIK:Xe+ +)°) =(Q IQ:
.Y(-, IIKII <, 11,
IIQII<1).
yf+ (X-) represents an operator ball in the Banach space (+Thus, ,-) ((-+ )) of all bounded operators T: ,Y+ - ,Y- (T: In what follows we shall often use the following well-known fact (see, e.g., [VII] ): ± The balls Jt are bicompact in the weak operator topology. )a E A of subsets of a certain set is said to be We recall that a system centralized if every finite sub-system of it 1_11 ;),"-, has a non-empty inter-
8.20
section n;,=1
11a, 96 0.
As is well-known (see, e.g., [VII]) Proposition 8.20 is equivalent to the following: 8.21
± Every centralized system (.X. kE A of sets .YCa E .Yf (a E A) which are
closed in the weak topology has a non-empty intersection.
Let 2+ (2-) be a non-negative (non-positive) lineal with the angular operator Ko (Qo), and let .X+ (Ko) (.W- (Qo) be the set of angular operators
1 The Geometry of Spaces with an Indefinite Metric
60
of all the subspaces from ff+ (elf-) which are extensions of '+ (2'- ), i.e., (cf. 8.7)
.X+(Ko)= (KI KE.X+,KJ Ko), .X-(Qo) _ (Q I QE.X , Q J Qo)
(8.20) (8.21)
In particular, if Y+ = (0) (Y_ = (0)), thenX, (Ko) =.X+ (.X_ (Qo) =.X-). We bring into consideration (see (8.9)) the sets Q*, QEX-(Qo)),
.X* (Qo)= (KI K= ,X+(Ko)= (QI Q=K*,KEX+ (Ko)). 8.22
(8.22) (8.23)
The set .X *_ (Qo) (respectively .X *+ (Ko)) coincides with the set of all
angular operators of subspaces from /lf+ (resp. from u-) which are J-orthogonal to the lineal _ (C' -) (resp. 2+ (D -O+ )) with the angular operator Qo (resp. Ko). The assertion follows directly from Theorems 8.11 and 8.11'. Theorem 8.23: The sets X + (K0), X- (Qo ) , X-* (Qo) and i and bicompact in the weak operator topology.
(Ko) are convex
We begin with .X+(Ko). Let K,, K2 E .X+(Ko). Since ,X+(Ko) C JY+, so by virtue of the convexity of .X+ when a E [0, 1 ] we have aKt + (1 - a)K2 E .X+ for any Ki, K2 E X+. Further, if x+ E cKo and K1, K2 E.X+(Ko), then
(aKi + (1 - a)K2)x+ = ceKtx+ + (1 - a)K2x+ = aKox + + (1 - a)Kox+ = Kox+, i.e., aKl + (1 - a)K2 D Ko, and the convexity of .X+(Ko) is proved. Now let a generalized sequence (K. hEr (C.X+(Ko)), where IF is a certain directed set (see [VII]) converge in the weak operator topology to a certain By virtue of 8.20 the ball .X + is closed in this operator k E topology, and so K E .X+. Since for any c > 0 and for any x± E ,Y± a yo E IF can
be found such that I ((K - K,,)x+, x-) I < e for all y > -yo, so we have
1((K-Ko)x+,x-)I 0, for all x+EVK.and forallx-E.W-; hence
( K - Ko)x+ = 0 (x+ E VK), i.e., K:) K o and K E X +(Ko).
Thus (see [VII]) X, (Ko) is a closed (in the weak operator topology) subset of the bicompact ball .X+ and therefore it is itself bicompact. The arguments are analogous for ,X- (Qo), X-* (Qo), and X+ (Ko).
Corollary 8.24: The sets X. (Ko)(1.X**(Qo) and x -(Qo)t1.X+(Ko) are convex and bicompact in the weak operator topology.
§8 The method of angular operators
61
Remark 8.25: In Corollary 8.24 it is not excluded, generally speaking, that the intersections mentioned in it may be empty. The non-emptiness of either of them implies, by 8.22, the relation .'+ [1] 2'_. Later (see Theorem 10.2) it will be proved that this is not only a necessary, but also a sufficient, condition for both the intersections mentioned in Corollary 8.24 to be non-empty.
Exercises and problems 1
Consider the two-dimensional real Krein space Yt' in Exercise 7 to §2 (see Fig. 1 there). In Fig. 2 is shown a non-negative (even positive) subspace 2'(C.Yt'). For any x E .P we have x = x' + x- (x± E Y ±), and x- = Kx', where K is the operator of rotating the vector x+ through an angle 42 (in the positive direction), and then
multiplying by the scalar k = tan p-the angular coefficient of the `line' Y. Compare this with the definition 8.1. Verify that II K II = I k I.
Fig. 2.
2
If in Exercise 1 p is always understood to be the minimal angle between 2' and 'the axis' .Ye', then tan w = II K II Prove that in the general case too (dim.Ye 500) for the angular operator K of a non-negative subspace 2' we have tan p(2', .YP') = II K II if
the (minimal) angle rp is defined by the equality sin w(2', .h° ) = suPe E s(r) II e - ZP' ell where S(2') is the unit sphere of the lineal 2' (II a II = 1) (M. Krein [XVII] ). 3
It follows from the result of Exercise 2 and Lemma 8.4 that uniform positivity of the subspace 2' is equivalent to the inequality p < u/4. Hence it is possible to derive in terms of the aperture 0(2',, 2'Z) = max(sin p(Yj,Y2), sin p(Yz, _Wj)) of two subspaces 2, and 2'2 (see, e.g., [I]) the following criterion: in order that all subspaces
.P' with a sufficiently small aperture O(2', 2') be non-negative it is necessary and sufficient that the subspace SP be uniformly positive (M. Krein [XVII] ).
1 The Geometry of Spaces with an Indefinite Metric
62 4
Let d' be a non-negative subspace with angular operator K and let 0 Or, JY ) < 1. Then 11 K11
6
`1+I.- IKIIZ
(Shlakman).
Hint: Use the properties of the aperture (see [I]) and the result of Exercise 2. 5
On the set of all semi-definite lineals 2' of a Krein space M we we introduce the functional
sup [e, e],
4'(l)=
if 2' is non-negative,
eES(i')
inf
[e, e],
if 2' is non-positive.
e E S(f' )
Suppose, for definiteness, that 2' is non-negative, and let K be its angular operator. It
is clear that 0 < c(.1) < 1 and that (c(2') = 01 a (91 is neutral). Prove that III K-1 II 2 1
-1
IIK
II
I
if K is bounded invertible,
+1
(Shlakman)
otherwise
1
Hint: By studying the function f(x) = (1 - x2)/(1 + x2) on the interval 0Sx K 11, convince yourself that 2
where µ = inf e'ES(P'Y')
1+/L
6
Ke+ .
Prove that a non-negative (non-positive) lineal 2' is neutral if and only if 4 (2',.x-) = I/ J (resp. 4 (2', . +) = l/J) (Shlakman). Hint: By comparing the results of Exercises 4 and 5 establish the formula
-1, (Y) = 1 -292(2, W-)(resp. 4?(2')= 1 7
Let 2' be an arbitrary subspace of the Krein space Jr = .Y+ ( .W-, and let 91 E d+ (see §8.8). If K is the angular operator of the lineal .', then
(K(*)x +X )x-EP-i'[0+]Ct' [(+]
(8.24)
and an analogous formula holds when 2'E,.d-. Here 9q' are the deficiency subspaces of the lineal 2' (Ritsner [4] ).
Hint. The `generalized adjoint operator' T(*) of a linear operator T. VT-,k GIT C .) should here be understood to be the operator whose graph I' T(*) is given by the formula F T(*' = ((u, v) E Wx f - I (Tx, u) = (x, v) for all X E V TJ (cf. (8.10)). For any bounded T such a T(*) exists and II T(*) II = II T 11. 8
Derive the formula (4.5) and Theorem 8.11 from the formula (8.24).
9
Suppose a linear operator E in the Krein space Jr = .r+ [(D] .e- satisfies the conditions EP+ = E, P+E= Then E is a projector (E2 = E), and further P+.
EW C P+ if and only if II Ell < J. An analogous result is true if P+ and J0 + are replaced by P- and -- respectively (Langer [2]). 10
11
Prove that the projector E in Exercise 9 and the angular operator K of the lineal E.,Y
(C .:O+) are connected by the formula E' = P+ + KP+ (in the case E.;( C .10- the corresponding formula is E= P- + QP-, where Q is the angular operator of the lineal E.) ([XVII] ). Prove that for the projector E in Exercise 9 the condition II E II < 2 is necessary and sufficient for the lineal Er to be uniformly definite. Hint. Use the result of Exercise 10 and Lemma 8.4.
§8 The method of angular operators 12
63
Prove that for a subspace Y, (EJ() of the class h' (see Definition 5.9) we have
Y1_= V,lEa/-flh-. 13
Let YE -I/-, let K be the corresponding angular operator, and let Pr, be the orthoprojector on to Y. Prove that (I+ + K*K) ' K(1+ + K*K)-'
P
(I+ + K*K) 'K* K(I+ + K*K)K II
(Azizov).
Hint: See the proof of Theorem 8.17. 14
Under the conditions of Exercise 13 prove that for the operator T= P+ I F: I - "Y' the adjoint operator 7-*: W+ ' has the form T* = T-' (I+ + K*K)-' (Azizov).
15
Under the conditions of Exercise 13 prove that the Gram operator Gv,(=P,JI ') has the form Gr = V*(I+ - K*K) V, where V. IF -i .YY+ is a linear homeomorphism, i.e., Gv- and (I+ - K*K) are `congruent' (Azizov). Hint: Use the results of Exercises 13 and 14.
16
Let _W- E .,//-, let Q be its angular operator, and let G,,- be the Gram operator for
99-. Then the operators Gr- and I- Q*Q are congruent, i.e., there is a linear homeomorphism V. 2W- such that Gi-= V*(I+ - Q*Q) V (Azizov). Hint: P- I F- can be taken as V. 17
Under the conditions of the preceding problem suppose infinite-dimensional uniformly definite subspaces. Then I K9,+ I = I+ + S
(I Kv-- I = I " + Si ),
+ (Y- _ / + [1]) has no
where S(Sj) E .gym.
(Azizov).
Hint: Use the results of the preceding exercise and of Exercise 7 in §6. 18
Let Y+ C .--O+ and Ky.. E y'm. Then ¶+ E h+ Q.W. Helton [2]).
19
Let V+ E .,//+ fl h+ and Y_ E <. //-. Then .Y'' fl y- = (0) a Y+ 4- 1- = .# (Azizov). Hint: Use Theorem 8.15.
20
Let .;f be a Krein space and let Y+ C '. Prove that 2' E..//+ if and only if 9'+[l] C.i°-. Hint: Use Theorem 8.11. '+ C YO C .I+ fl Y+ [1] , and .1= Yd1J IIo. Prove that I+ E J(+ (.y') if and only if T+/moo E ,//+ (. ). Hint: Use the definition of the factor space . and Exercise 20.
21
Suppose .W is a Krein space,
22
Suppose
is a Krein s+pace, Y± C e°±, and Yo c Y'± fle°°. Prove that, if
dim Fo < co, then 9+ E h- if and only if t'±/go E h± (Azizov). Hint: Verify that the subspaces I'± E h- if and only if their angular operators can be expressed in the form of a direct sum of a finite-dimensional isometric operator and a uniform contraction, i.e., an operator T with II T II < 1.
23 We introduce on I(+
a
functional
0(.IF,.,t) = sup (O(x, y) I x E ', y E .,I
Y', 't"E elf+), where
O(x,y)=1 if Re2[x,y] _< [x,x][y,y];
B(x,y)=252+26,62-1-1 if Re2[x,y]> [x,x][y,y]>0 and 62 = Re2 [x, y]/ [x, x] [y, y]; 0(x, y) = 0 in the remaining cases. Prove that the relation Y- .4 which is expressed by the inequality 0(s(1, .4 ") < oo is
1 The Geometry of Spaces with an Indefinite Metric
64
W+, that on each equivalence class the function an equivalence relation on p(/,.T) = In B(!,A") is a metric, and that each equivalence class is a complete metric space (A. V. Sobolev and V. A. Khatskevich, [1], [2]).
24
Prove that the set /to of maximal uniformly positive subspaces in an equivalence class (Khatskevich, [13], [14]).
25
Using the connection between operators from tl+ introduce an equivalence relation on
and subspaces from
K,-K2 ifO(Y,,.'2)<-, where 2',=[x++K,x+Ix+E.
+J,
i=1,2.
Verify that the interior of the ball .X+ (W+, .W ) is an equivalence class (Khatskevich, [13], [14]).
§9
1
Pontryagin spaces H,r. W('r)-spaces and G('r)-spaces One subclass of Krein spaces is particularly important in operator theory.
Definition 9.1: A Krein space
= M+ [+]
with a finite rank of
indefiniteness
K =min [dim
.W+, dim W-) (see Definition 6.5)
is called a Pontryagin space and is denoted by H. Admitting a certain freedom
of speech, we shall sometimes call x the number of positive or negative squares.
In accordance with this definition we shall denote a fixed canonical decomposition of a Pontryagin space Hx by IIx = H+ [+] II
(9.1)
and without loss of generality we shall in future assume, if nothing to the contrary is stated, that (oo > )x = dim II+. Otherwise we would have changed over to the anti-space (see Definition 1.8). Thus the equality x = 0 corresponds to the (trivial) case when IIx = TI- is a Hilbert space with the scalar product (x, y) = - [x, y] (x, y E IIx) (see Remark 2.3). Examples of spaces IIx are given in the Exercises for this section.
It follows from Definition 9.1 that all the general geometric facts about Krein spaces are true for a space H. However, and additional condition, namely, the inequality x < co naturally gives rise to a number of additional properties of Pontryagin spaces. It is with these properties that we shall concern ourselves. The simplest of them is contained in the following proposition. 9.2 All non-negative lineals of the II, are finite-dimensional.
By Corollary 4.3 we have for any non-negative 9 (CI-Ix) that dim Y < x(
§9 Pontryagin spaces II.. Wax'-spaces and G(x)- spaces Corollary 9.3:
65
In the space II, every positive lineal is uniformly positive.
This follows from 9.2 and 5.4.
9.4 A non-negative (positive, neutral) lineal 2' in the space IIx is a maximal non-negative (a maximal positive, a maximal neutral) lineal if and only if dim Y= x. This follows from Corollaries 4.6 and 4.8 and Theorem 4.13 (taking 9.2 into account). In applying Corollary 4.8 it should be borne in mind, that in contrast to the general case, in the space Hx all maximal positive lineals are closed (because they are finite-dimensional). To conclude this section we establish a lemma which will often be applied in the theory of operators acting in the space II{. Lemma 9.5: Any lineal 2 dense in IIx (2'= II.) contains a maximal (i.e. a x -dimensional) positive subspace.
We consider the canonical decomposition (9.1), where dim II+ = x, and a ..., ex) in II+. It is clear that [ei, ek] = (ei, ek) = bik, (j, k = 1, 2, . . ., x). For any c > 0, because of the condition 2 =1I,,, vectors f l , f2,. ., fx E 2' can be found with the property I f k - ek 11 < (k = 1, 2, ..., x). Since all the Gram determinants of the system {el, e2, ..., ex) are positive:
certain orthonormalized basis (el, e2,
detll[ei,ek]IIi"k=1=detllbikII 'k=1=1 > 0
(m=1,2,...,x),
it follows that, for sufficiently small e > 0, all the determinants det II [fi, fk) IIl k=1 (m = 1, 2,
.
.
.,
x)
are also positive,
i.e., the Gram matrix II [fi, Al II k=1 of the subspace . = Lin(fl, f2, ..., fx] (C 2') is positive definite: thus the vectors fl, f2, . . ., fx are linearly independent, and 9 is a x-dimensional positive subspace.
The condition x < co, apart from its natural influence (as explained in section 9.1) on the character of non-negative lineals (subspace) of the space IIx, also has an influence on the negative subspaces of jr,,. 2
Theorem 9.6:
All negative subspaces 2' of the space IIx are uniformly
negative.
By virtue of the obvious analogue 8.9' of Remark 8.9 there is a maximal negative subspace -Tmax J Y. By Corollary 8.131, 2ma1x is a maximal positive
subspace and, more than that (see Proposition 9.2 and Corollary 9.3), it is a uniformly positive x-dimensional subspace. By virtue of Corollary 7.15 it is projectively complete, and since (see Proposition 7.9) 2'm.., = is also
66
1 The Geometry of Spaces with an Indefinite Metric
projectively complete, it is also uniformly negative. Therefore (see 5.3) 2' too is uniformly negative. Corollary 9.7:
All definite subspaces 2' of II, are projectively complete.
This follows from Corollary 7.15 and Corollary 9.3 (for positive 91) and from Theorem 9.6 (for negative 2'). Remark 9.8: The assertion of Theorem 9.6 cannot by any means be extended to non-closed negative lineals. The example 4.12, which we have already used more than once, bears witness to this (with accuracy up to a transition to the anti-space). The space W considered in this example is (up to a sign, i.e., up to a transtion to the anti-space) the space II,, and the non-closed positive lineal 2' constructed in it is not uniformly positive (as was explained later in Example
5.10), i.e., 2' is singular (cf. Exercise 6 to §5). We shall return later, in Exercises 9-12, to the analysis of similar situations. Theorem 9.9: In a space A. a subspace 2' (=.') is projectively complete if and only if it is non-degenerate.
We recall (Proposition 7.10) that non-degeneracy of a subspace 2' in the
general case of a Krein space is a necessary condition for its projective completeness. We show that in our case it is also sufficient. If 2' is non-degenerate, then in any canonical decomposition of it . = 2+ [ 4- ] 2(which exists by virtue of Theorem 6.4) the definite subspaces Y+ and 2- are uniformly definite (Corollary 9.3 and Theorem 9.6), and therefore (Theorems 7.16 and 5.7) 99 is projectively complete.
Corollary9.10:
Every non-degenerate subspace Yin n. is itself a Pontryagin
space II,, with a certain rank of indefiniteness x': 0 < x' < x. 2' is projectively complete, and in its canonical decomposition 2' = 2+ [+] 2-, by virtue of Proposition 9.2 0 < dim 2+ < x, so that when dim Y+ < dim 1- we have x' = dim Y+ < x, and when dim 2- < dim 2+ a
fortiori x' = dim 2- < x.
Comparing some of the facts discovered in §9.2 we remark that they are characteristic for picking our Pontryagin spaces in the class of all Krein spaces. 3
Theorem 9.11: For a Krein space .Y( = W+ [+] . the following three assertions are equivalent: a) the space .
is a Pontryagin space (,W = n,);
§9 Pontryagin spaces II,,. W'")-spaces and G(")- spaces
67
b) all the definite subspaces 2' (CM) are uniformly definite (projectively complete); c) all the non-degenerate subspaces 2' (C,W) are projectively complete.
That a) - b) - c) was established in §9.2. It remains to prove the implication c) - a). To do this it is sufficient to adduce an example of a non-degenerate but not projectively complete subspace 2' in any Krein space of infinite rank of indefiniteness: x = min (dim ,+, _+ [+] dim ,-) = co. But as such an example any subspace 2' which is, say, positive but not uniformly positive will serve (see Exercise 1 to §5).
In studying the spaces II" (and later the operators acting in them), instead of the prefix 'J-' (J-orthogonality, J-isometry, etc.) which is traditional for 4
Krein spaces, the prefix '7r-' or `a"-' is more expressive since it indicates at once
that everything takes place in a Pontryagin space, and the second form indicates the rank of indefiniteness. In particular, we shall in future use these symbols, naming, for example, vectors x, y (E II") for which x[1] y, and sets 9', ,tf (C IZ") for which 2' [1] /u as a-orthogonal, and we shall call the lineal -1t['] the 7r-orthogonal complement of the set elf. We recall (see Proposition 9.2) that the isometric lineal 22° = 2' n 2' of a
lineal 2' in a space II" is always finite-dimensional (dim 2° < x). For degenerate subspaces 2 (2= 2', 22° = (B)) in II" instead of Theorem 9.9 we have to be content with a more complicated decomposition (generated by 2')
of H. into a direct sum of subspaces more complicated than the simple a-orthogonal sum 2 [+] Y111.
Theorem 9.12: Let 2 be a subspace in II,,, let . tf = 2' be its 7r-orthogonal complement, and 22° = 9 n , It be the isotropic part of Y. Then the following decomposition holds: H. = 21 [+] -MI [+] (2° + 4'),
(9.2)
where 2, and itf, are non-degenerate subspaces connected with .9' and It respectively by the relations
2' = 2, [+] 2'°,
'tt = -it, [+] 22°,
(9.3)
and .4 is a certain subspace skewly linked with 2'° (see Definition 1.29): .4 - # 22°. For any choice of 2, and tf1 in (9.3) the subspace 1'skewly linked with 22° may be chosen arbitrarily in the subspace (2, [+] On the other hand, for any choice in II" of a subspace .41" (#22°) satisfying the conditions (9.3) the subspaces 2, and .,tl, in (9.2) are defined uniquely: _r, = 2' n 1"[1],
at, =,m n X1.111.
68
1 The Geometry of Spaces with an Indefinite Metric
O The subspaces 21 and .,lt, satisfying the conditions (9.3) (by Proposition 7.3 2° is the common isotropic subspace for 2' and .fit) are not-degenerate, and therefore (Theorem 9.1 lc) they are projectively complete, and the same, as
is easy to understand (see Exercise 8 to §7), applies also to their
ir-orthogonal sum 2t [+], th (we recall that 2t C 2 and Wl C . tt = Y[1). are certain Pontryagin spaces (see Corollary 9.10). The relations (9.3) show that the second of them contains 2'° Thus, both 2't [ + ] , f 1 and (91 [ 4 ] I ( )
-I
(dim 2'° S x). As in every Krein space, in the space (21 [-+] .A(l) [1] there are (see Exercise 4 to §3) subspaces skewly linked with 2'°. We choose any one of them and call it
_f': A' # 2?°. Then by virtue of Lemma 1.31 and Proposition 1.30 dim
.4'= dim 20 and the direct sum 2° + N is non-degenerate, i.e.,
it is
projectively complete. Thus (see Exercise 8 to §7) the subspace 3 _ 21 [+] J11 [+] (2° +,A,)
is also projectively complete, i.e., it is non-degenerate. Now let x be such that x [1] R. In particular, x E (2'° + 21) [1] = 2[1] and X E (2'° + fil )[1] = -0[1] = 27[1] [1] = 2, i.e., X E 2' n 2[1] = 2'°. On the other hand, x [1] .,N and therefore, since .4' # "?°, we have by virtue of Definition 1.2 that x = 0. Thus, R = IIX and the equalities (9.3) are proved. We pass on to the last assertion of the theorem. We fix in IIx an arbitrary V
(# .'°) and we put Yl = 2' n .N[1] . The lineal 2t is obviously closed; 2'° [1] 9-91 (since 2'° [1] 2') and -wo n gel = (6] since 2'° # N. Further, 2° + N[1] = II (see Corollary 1.32), 2° C 9, and therefore (YO +,4-1-0) =YO+(yn,101-1)= YO
Similarly one verifies that for tt l = ff n 1011 the relation A( = 2'° [+] J11 holds. Thus, for the constructed subspaces 2t and ,A11 the relations (9.3) hold and T1 n (6]. Since by construction 2'(1] = (2 n IV[1])[1] D .,1 [1] [1] = .4-
and
similarly
,
&{1] = (.,lt n .ii [1]) [1] D 4-,
so
.'i" C (91 + 41l)[], and by virtue of the first part (already proved) of the theorem the decomposition (9.2) holds.
We now prove that, in the decomposition (9.2) just constructed, the subspaces 2l and -01 satisfying the conditions (9.3) cannot, for a fixed N (# 2o), be chosen differently. For, if 2t and -01 satisfy the conditions (9.2) and (9.3), then 21 C 2 n.-Ii'[-L], -itl C at n ./i"[1], (9.4) However, we saw earlier that the subspaces 91 n ,4'[1] and -it n .4"[1] also satisfy the conditions of the form (9.3): Se = (2' n,/1-111) [+] 9'°, !1= ( It n . i "[1]) [+]°; therefore it follows from (9.4) and 1.21 that '1 = 2 n .'V[1] and at, = at n ..d'[1].
5 We present one more simple proposition, which is often applicable in investigations.
§9 Pontryagin spaces II,,. W')-spaces and G(")- spaces
69
9.13 If .' (C II") is a space with inertia index (7r, v, W), and if 2° is its isotropic subspace (dim 2° = w), then the factor space 2/2o with the
indefinite metric induced from 9 according to the rule (1.16) is a Pontryagin
space r, where x' = min (7r, P). We express 9 in the form 9 = X, [+] 27°, where 91 is a non-degenerate
subspace. By virtue of Corollary 9.10 9, is a Pontryagin space II, :
a canonical decomposition of 2, have the form where, clearly, min (dim 9i ,dim 9i )= x' . But Y _ 271+ [+] 9i [+] 9° is the canonical decomposition of the whole of 2, and so x' = min{7r, v} (Theorem 6.7). It remains to apply 1.23.
0<,x' <,x.
Let
Corollary 9.14:
Let 90 be a neutral linear in II", dim 2'° = w ( x), and
9 _ (270)[1] . Then 92'° is a space II,,, with x' = x - w.
Since 2° is an isotropic subspace for 27 (see Proposition 7.5), so In 9 = (jr, v, w} with certain 7r and v, and therefore by virtue of Proposition 9.13 2/2o = II,,-, where x' = mina, v). But it is easy to see that here a = dim II+ - w, because 9 J V ° (the deficiency subspace for 2°) and dim 9 ° = x - w, and v = dim II-, so that v > x > x - w = a and therefore
x' =x-w. 6 Among W-spaces (see §6.6) and in particular G-spaces the analogues and generalizations of Pontryagin spaces II" are of particular interest in the theory of operators. We shall denote by x the maximal dimension of the non-negative subspaces contained in a W-space W. If x < oo we shall call ,Y a W")-space, and if it is non-degenerate a G(")-space.
9.15 A W-space .3f is a W")-space if and only if the whole positive part of the spectrum a(W) of the Gram operator W consists of eigenvalues and the
sum of the multiplicities of all non-negative eigenvalues is equal to x. Moreover, ,;t' =.W/Ker W is a G("-"°)-space, where xo = dim Ker W.
If the spectrum a(W) satisfies the requirement indicated, then (cf. Theorem 6.4), in the canonical decomposition
.re=,w-[G) ].WO[G) ].Y+
(X:2:
defined by the spectral decomposition W= f aid of the projectors
,.fo=P°.() =P+-
(9.5)
X dE,, (E),+o = Ex) with the rr
P-= f _
dE>,, m
Po=Eo-E_o,
P+=
JMdE. 0
the (isotropic) subspaces Y° (= Ker W) is neutral, and also (see Proposition 6.5) dim X0 = xo is the multiplicity of the number 0 E Op( W), and the maximal dimension of the non-negative subspaces contained in .W' - [ O+ ] ,Y+ is by Lemma 1.27 equal to dim + = x, the sum of the multiplicities of the positive
1 The Geometry of Spaces with an Indefinite Metric
70
eigenvalues of the operator W. Thus the maximal dimensionality of the non-negative subspaces in is equal to xo + x, = x. Conversely, if . is a W(')-space, then in its canonical decomposition (9.5) dim(, W' [ (D].+) = x (again by virtue of Proposition 6.5 and Lemma 1.27).
The last assertion regarding .' follows from Proposition 1.23. Corollary 9.16: A G-space ,Y is a H. space if and only if it is regular and dim 2' (G) = x < oo, where 2+ (G) is the direct (and orthogonal) sum of all the eigen-subspaces of operator G corresponding to its positive eigen values. If .7" is a W(`)-space, then . = lKer W(') is a IIx xo space (xo = dim Ker Wtx)) if and only if . wo) = WO).
The condition for regularity (0 E p(G)) is connected with the fact that a IIx space is a Krein space (see Proposition 6.13), and all the rest-follows from Proposition 6.13. But if a G-space a e is singular and is a G(')-space, then it can be `densely imbedded' in a II, space by means of the procedure described in Proposition 6.14.
To conclude this paragraph 6.9 we indicate another useful criterion for a G-space to be a G(x)-space.
9.17 Let = _T+ V be a G-space, 2 and N subspaces of it with 2' C +, dim 2?= x < oo, and iV C #I-- U (9). Then Yeis a G(')-space. It follows from Proposition 1.25 that 2' is a maximal non-negative subspace. It remains to verify that .
does not contain non-negative subspaces
2' of higher dimension than x = dim Y. Let P+ and P- (P+ + P- = I) be the projectors generated by the direct decomposition . = 2 + N, and let 2' C ,0+. Then by Lemma 1.27 the mapping P+: 2' - 2' is injective, hence
dim 2' < x.
Exercises and problems 1
Let
A=
be a linear space of finite sequences x = (i )7=, (E; = 0, j > Ni), and let aik II7k., be an infinite Hermitian matrix defining on. an Hermitian metric
[x, y] = Z aik jI k i.k=,
(x
y = (nl k=, E .f ).
Let the matrix A be such that i does not contain isotropic vectors (J ° = (0)), and each of the forms [x, x] = E k (x E 9) contains not more than x, and at least one, positive squares. Prove that . after completion (cf. Exercise 4 to §2) becomes a rL, space ([XIV] ). 2
Prove that if the condition on the number of positive squares for the forms in Exercise 1 is observed, then the condition for non-degeneracy of the metric [x, y] on .(i.e. Jr .Y° = (0)) is equivalent to the requirement det 11 aik IL k=, ;4 0 for all sufficiently large n. Without the condition on the number of positive squares, the given requirement is
§9 Pontryagin spaces IL. W"-spaces and
spaces
71
sufficient, but not necessary, for the non-degeneracy of the form [x, y] (I, Iokhvidov [19]). 3
Modifying the example in Exercise 1, consider the linear space Jr of two-sided and finite (on both sides) sequences x= ( . . . , i;-,, to, ti, E2,. ..), and an infinite (in all four directions) Hermitian matrix A = I I a i k II k=-m; formulate the corresponding conditions that will ensure that . (after completion) becomes a IIx space ([XV], cf. [V] ).
4
Verify that in the examples in Exercises 2 and 3, if non-degeneracy of the metric [x, y] is not demanded (i.e., if To (B) is allowed), then the factor space after
completion becomes a H. space ([XV], [V]). Hint: Compare with Proposition 9.13. 5
Let JP= 12 be the space of all sequences x= (SxJk=,, k=, 12 < 00; the form [x, y] for x, y E 12, y = (tlkI k=, is given by the formula [X, Y1 _
Zkx
=1
r
t
Zk=x+1 Sk1jk
Compare this with the example in Exercise 4 to §2, and show that .' is a IT. space (Pontryagin [1]). 6
Let . be an abstract set on which a complex Hermitian kernel K(p, q) = K(q, p) (p, q E . ) is defined such that the form E k=, K(pi, pk)(jEk for sufficiently large n
contains exactly x positive squares. The lineal consisting of all functions of the form 'P(p) =
j=1
IjK(p, qi),
(p) =
j=1
?jK(p, qi)
with the form [1p,,1] = E k=I K(qi, gk)Ej k becomes, after taking the factor-space by the isotropic lineal and completion, a IT, space; this is the continual analogue of the space in Exercise 1 (M. Krein; example published in I. Iokhvidov's report [4] ). 7
In the example in Exercise 6 to §2 the space L2(S, E, µ) will be a IIx space if in the canonic Jordan decomposition it = µ+ - µ- of the measure into the difference of two
positive measures the component µ+ is concentrated in x points of the set S not belonging to the carrier of the measure µ_ (see I. Iokhvidov [2] and Azizov [8], where particular cases of this situation, analogous to the example in Exercise 5 to §2, were considered, and also [IV] ). 8
In the real space II, for any two non-negative vectors x, y (EII1) establish the inequality [x, y] 2 > [x, x] [ y, y], in which the inequality sign holds only when the vectors x and y are collinear (M. Krein, see [XV], see also M. Krein and Rutman [1]).
9
Consider IT. = II+ O+ II-, where 0 < dim II- = x < co, and II+ is infinite dimensional. Let yo E IIx and let . be a positive lineal in II.. Then for continuity of the Iv it is necessary functional p,.,(x) = [x, yo] (xE 91) relative to the internal norm and sufficient that yo [1] (2)° (I. Iokhvidov [11]). Hint. In the proof use the canonical decomposition of the closure 2': 2' = Y', + (2')° and the fact that a positive subspace 2, is projectively complete (Theorem 9.11,b).
10 A definite lineal 2' (C IIx) is singular if and only if its closure 2' is degenerate (I. Iokhvidov [11]; [VIII]). Hint: Use the result of Exercise 9. 11
Let 2' (C IIx) be a singular lineal. Then for the equality 2' = 2' [+] (2')° it is necessary
that 2' should be complete (i.e., a Hilbert) space relative to the internal metric (I. Iokhurdov [11]). Hint. Use Proposition 9.13.
1 The Geometry of Spaces with an Indefinite Metric
72
The condition of completeness of ' with respect to the internal norm
12
z- is
sufficient for the representation .e = . [+] (l )° of the closure of a singular lineal 99 in any Krein space . (I. Iokhvidov [I fl). Hint: Prove that the lineal' [+] (2')° is closed relative to
Consider the Pontryagin space rl, = fl+ [ (@] II- (dim IV < oo ). Prove that a
13
sequence
strongly converges to a vector xo when n - oo if [x,,, x] - [xo, xo] and
[x,,, y] -' [xo, y] for y running through a set 1 dense in fl, (M. Krein and Langer [3]).
Hint: Without loss of generality it may be supposed that 1 is a lineal and fl+ C 1. Verify first that x - xo, and then use the ordinary criterion for strong convergence
for [x,-)i. Let .', = W' [ Q+ ] .Y; be a J-space, and let .02 =
14
z+ [ O+ ] .W 1 [ O+ ] 7Yz be the
canonical decomposition of a W-space . '2, where W= T" JT, and T:.Y°2 - .', is a bounded operator. Then dim .YZ < dim W' , and therefore, in particular, if .,Y, is a Pontryagin space with x positive (negative) squares, then the positive (negative) part of the spectrum of the operator W consists of not more than x (taking multiplicity into account) eigenvalues (Azizov).
§10
Dual pairs. J-orthonormalized systems and bases
Definition 10.1: A pair of subspaces
1
of a Krein space
is called a dual pair if 2? ± C .? ± and 2'+ [1] 2_ . If, in addition, 2± E tl±, then (2+, y2_) is called a maximal dual pair. A dual pair (2+, 2-) is called an extension of the dual pair (2+, 2_) if 9?± D Y+, and it is called a maximal extension of the pair (2+, 2_) if this extension (2+, 2-) is a maximal dual pair. Theorem 10.2: (7max omax)
Every dual pair (2+, 2-) has at least one maximal extension
Moreover, every maximal in 2+11 non positive extension 2_ of the subspace
2_ belongs to 4Y- and the pairs
(2[13, SP_) exhaust all the maximal
extensions of the dual pair (2+, 2'_ ). Clearly it suffices to prove the second assertion of the theorem. Consider
the sub-space X+11 (32_ ). Let w y+ _ )1+ = (P+2+)1 fl T+ be the deficiency subspace for 2+ in X+. We recall (see the proof of Theorem 4.5) that !2+ [1] 2+ and Therefore 2+ [+] 9+ E -It+. 2+ [+] 1i+)[11 E , _' (see Theorem 8.11) and . V = 4"0 [ (@].IV1, where .4"° is the isotropic subspace for V, and ./{", is the negative subspace. Since A{" C YA+1 and Cl+ C 2+1], it is clear that Y+11 J .4-0 [ O+
O+ ] `c+
Conversely, if z [1] 2+, then z = z+ + z- (z` E .Y±), and z+ = zi + z2+, where z1 E 1+, and zz E (g+ )111 n o+ = P+2+. Thus z = zi + (zz + z-), and, at zz + z- = z - zi [I] 2+, the same time, zz + z E 2+1 fl (go+)111 = (2+ [+] Qo+)111 =,4 (see (1.12)), and so i.e.,
§10 Dual pairs. J-orthonormalized systems and bases
73
z = (zz + z-) + z1 E l' [ ] 91+. Finally we obtain for Y+1] the canonical decomposition rte[+1
=
g+
[O] X1"1
-0.
(10.1)
Now let be a maximal in 94+l non-positive extension of the subspace '_ (C + 1). Then (see Proposition 1.22) 2_ D N° and Je_ = 4 -0 [ Q+ ],JP , where
Y_ is the maximal, in the subspace 9+ [ @]./V1, non-positive subspace (see (10.1)). We have turned out to be in the situation considered in Proposition
8.18, by virtue of which 2_ = (Fx+x),XE.,i where the linear operator F: ,A", - !2+ is defined on the whole of ,4', in view of the maximal non-positivity in g+ [ @] 4'1. Therefore of
P - = Lin(P-,N°, P _) = Lin(P-,IV°, P-.4 1 J = P- (,/V° [ G) ]. {-1) i.e., k_ E .elf-, and
is one of the extensions of the dual pair (Y+, '-) of
the form (27max' max)
The last assertion of the theorem is obvious. Definition 10.3: Let W be a Krein space, let be a maximal uniformly definite dual pair (i.e., 2'± are uniformly definite), and let 9f be a 2
subspace in f. We call the decomposition 9 = 1o [+] 91 [+] V2 of the
9
(2'+, 2'_ )-decomposition if go = 91 fl g 111, the orthogonal complement to c 0 [+] 911 relative to the scalar product (x, y)1 = [x+, y+] - [x_, y-], where x = x+ + x_, y = y+ + y_, x±, y± E Y+, and therefore also relative to the form [x, y] (see Exercise 8 to §2).
subspace
its
1 = ( fl +) [+] ( fl _ ), and g2 is
We note that any subspace CA admits an (2'+, 2'_ )-decomposition relative to any uniformly definite dual pair and this decomposition is unique.
Theorem 10.4:
Let 9) = 910 [+] 91 F+1!22 be the (.'+, Y-)-decomposition of a
subspace !2, and let qi2 = g2' [+] 92", where 92' is a projectively complete subspace. Then there is a maximal uniformly definite dual pair (2'i , 9,1) such that the decomposition 9) = go [+] (ci [+] i) [+] 92 of the subspace ft is its (91', 2'i )-decomposition. Let 92'= 9+' [+] 9- be a canonical decomposition. We prove our assertion for the case 92' = 9+ and by so doing reduce it to the case 92' = V-' which is treated according to the same scheme. Thus, 92'= g+. Without loss of generality we shall suppose that 0± = Y±, and we put 2'i = 1'2' [+] w, fl 912' I']). By construction 2'1+ is a uniformly definite subspace and, 'z fl 2'i = (0). Moreover, Y1 E ,ff +; for if this were not so, then in + there would be a vector xo ;e 0 J-orthogonal to 2' and therefore to C42', i.e., xo would lie in .,Y+ n92' [ll and therefore (xo , xo) = 0; we would have a contradiction.
1 The Geometry of Spaces with an Indefinite Metric )(1] Then by (1.12) -ri = (.W- + 1z) n vz We put 21 = Since V2' C 9, and v n ._w+- C v 1 C v2, [1', so v n 991, = Sz [+] (1 n -w+ ) and v n &i = g n [(. - + g2) n v2 [1]) = v n.-*,-, i.e., (, n Yt) [+] (g n -ri) = 91 [_-] c'2. Hence c'2,, n LI = (B). It therefore 74
remains to verify that V2 is orthogonal to Vo [+] 11 [+] 92 in the scalar product (x, y)1 indicated in Definition 10.3.
Let x E gz . In accordance with Exercise 8 to §2 (x, y)1 = 0 when y E 21 [+] 92', and therefore the canonical projections xi of the vector x to Pi lie in . ± n cz [1] respectively. Consequently C (.4+ n!22, [1]) [+] (,Ye- n C/Z [1]) _ 1. From Exercise 9 to §2 we conz
on
clude that go C .1. It remains to notice that the scalar products (x, y)1 and
(x, y) coincide on MI, and therefore also that (x, z)1 = (x, z) = 0 when x E 92", _ o [+] (°11 [+] I Z) [+] I z is an z E 'o, i.e., the decomposition
(Let, Li )-decomposition of the subspace V. 3
Later in Chapter 3 we shall encounter more than once dual pairs and
maximal dual pairs of subspaces invariant relative to one or other operators or
even whole families of operators. But in this section dual pairs arise in a natural way in the consideration of the so-called J-orthonormalized systems in a Krein space.
Definition 10.5: A system of vectors 9= (ea )a E A (CO), where A is an arbitrary set of indices, is called a J-orthonormalized system if [e,,,eat] _ ±Sa,a, for all al, a2 E A, where Sa,a2 is the Kronecker delta. The simplest example of a J-orthonormalized system of Y = + [ ] .'is the union of two arbitrary orthonormalized (in the usual sense) systems from
the subspaces -W+ and- respectively.
10.6 Any J-orthonormalized system j= (e, )a E A is algebraically free, i.e., it does not contain linearly dependent finite-dimensional sub-systems;
moreover, it is minimal in the sense that not one vector ep E . belongs to C Lin(ea)aE A, , 0
.
It is clear that the first assertion follows from the second. But the second is we have obvious, since (see Proposition 3.5) for any x E C Lin (ea )a E A,,,
[x,eQ]=0 and [ea,e0]=±1. Definition 10. 7:
A J-orthonormalized system is said to be maximal if it is not
contained in any wider J-orthonormalized system, and to be J-complete if there is no non-zero vector J-orthonormalized to this system (see Remark 10.11 below).
10.8 Any J-orthonormalized system is contained in a certain maximal J-orthonormalized system.
§10 Dual pairs. J-orthonormalized systems and bases
75
The proof is obtained in the usual way using Zorn's lemma (see Theorem 1.19).
Every J-orthonormalized system [ ea )a E A in a Krein space .0 generates a dual pair of subspaces (2'+, 9?-), where 4
_T+ = C Lin ii+,
1 + = [e, I [ea, a«] =± 11aE A.
(10.2)
10.9 If a dual pair (2'+, 9-) is generated by a J-complete J-orthonormalized system 9 (see (10.2)), then the subspaces 2'+ are definite. It must be shown that 2+ are non-degenerate. But if an isotropic vector were found in _T+ (or 2-) is would be J-orthogonal to the whole system ., which contradicts the J-completeness of J. At the same time maximality, or even J-completeness, of the system 9 does not, generally speaking, imply the maximality of the corresponding dual pair. What is more, the following proposition holds: 10.10
In any Krein space W distinct from H there are J-complete J-
orthonormalized which generate dual pairs (2+, 2'_) that are not maximal dual pairs. We choose in .
an arbitrary positive but singular, i.e., not a uniformly positive, subspace Y+ E . It+ with the angular operator K, 11 K 11 = 1 (see Lemma 8.4). The corresponding Gram operator G,,, is invertible (0 ¢ ap(Gv,)), but Gv,,' is not bounded and defined on a domain ?c,_ (# 2+) dense in 2+ (see Propositions 6.5 and 6.11). We consider an arbitrary vector xo E 2'+\3G.,.. and the subspace 9+ = se+ e Lin[xo]. We notice at once that Z;7,Y, = Y+, for it would follow from zo E 9?+ and zo 1. G,,. ` +'
that (Gv-+zo, v) = (zo, G,- v) = 0 for all u E Y+, i.e., G,-+zo = Xxo, and since xo¢WG,_, so X=0, and hence Gv-, =0 and zo=0. Now in the space 2+ we choose any complete orthonormalized (relative to the scalar product ( , )) system (a Hilbert basis), and we orthonormalize it by the Schmidt process relative to the intrinsic metric [, ] which is positive on the whole of 2+ and, in particular, on 2+. It is asserted that the Jorthonormalized system . + so obtained, for which obviously C Lin.J+ = 9'+, is J-complete and therefore maximal in -T+. For, if a vector yo E _T+ could be found such that yo [I] . +, then (Yo, Gg,, u) = (Gv-, yo, y) = [yo, u] = 0 for all u E . + and and yo = 0. yo [1J Gv-+ (C Lin . + = Gr,9' , i.e., yo 1 Thus, C Lin + = .T+ and 2+ 0 X11+. Now it remains to choose in the subspace Y_ = 2'+1' (E..1(-) an arbitrary complete orthonormalized system and reconstruct it by means of the Schmidt process into a J-orthogonal system
.J -. The union
_+ U. _
is a J-orthonormalized system, and is also
1 The Geometry of Spaces with an Indefinite Metric
76
J-complete, for
x(J-1 J = x [L] J_ = x[l] C Lin .
_
i.e., x E -T+.
But x [1] . _ , from which it follows that x = 0. At the same time the dual pair
(M+, 9-) generated by the system is not maximal. Remark 10.11:
In the course of the proof we have incidentally discovered the
difference (for subspaces) between the concept of the J-completeness of a system and its completeness in the usual (Hilbert) sense: the system J+ we constructed at the beginning of the proof is J-complete in the subspace 9+ but is by no means complete in it: 0 ;4 xo E 9+ O C Lin J+. This is explained in the present case by the singularity of the subspace 2+, regarding which we spoke (allowing a certain freedom of speech and interpreting Definition 10.7 broadly) of
the J-completeness and completeness of the system J, But as regards a J-orthonormalized system in the whole Krein space it is obvious that its J-completeness and ordinary completeness are equivalent.
5 We now consider J-orthonormalized bases in a separable Krein space .,Y = ,e+ [ O+ ] W-, i.e., J-orthonormalized systems which are (Schauder)
bases in r. 10.12
If a J-orthonormalized system Jr = [ ek ) 'k= I is a basis in .W, then the
pair (9+, T_) generated by it is a maximal dual pair.
E We prove, for example, the maximality of T+. Let us assume the contrary: 9+ 0 + Then P+24 # W+, and in the deficiency space 9+] (1 j+ there is a non-zero vector xo = Ek=1 kek. Let (eki] 1 be the set of all vectors of the basis (ek]k=I for which [ek, ek] = 1. Then 0 = [xo, eki] = tki (i = 1, 2, ...), from which [xo, xo] 5 0, i.e., xo = 0, contrary to hypothesis.
We recall that a system of vectors (fk)k=1 in a Hilbert space is said to be normalized if II fk II = 1 (k = 1, 2, ...) and almost normalized if there are positive constants µ and M such that (0< )µ S I I fk I I < M (k = 1, 2, ...). The systems (fk)k=1 and (gk]k=1 are said to be biorthogonal if (fj, gk) = Sjk (j, k = 1, 2, ...). It is well-known (see, e.g., [XI]) that every system biorthogonal to a normalized system is almost normalized. 10.13 Every J-orthonormalized basis is almost normalized. Let (ek) k= I be a J-orthonormalized basis. Then
the system (e°k = II ek II-1ek)k=1 is normalized, and, as may easily be verified, the system { hk = sign [ek, ek] II ek I I Jek) k=1 is biorthogonal to it, and is therefore almost normalized: 0 <,u <, hk I I < M (k = 1, 2, ...). But since 2 II ek II_ek = sign [ek, ek] Jhk, so and therefore II ek II = II Jhk II = II hk II ,
0 <,µS IIek11<, M.
§10 Dual pairs. J-orthonormalized systems and bases
77
Exercises and problems Let (.P+, 9'-) be a dual pair of subspaces, let K and Q be the corresponding angular
1
operators, and let 1'7, be the deficiency subspaces for Y± respectively (see §6.1, §6.2). Then
Kx=Q(*)x+ W+(I+ - QQ(*))"2x for all XE P+2+, Qy= K(*)y+ W-(I- - KK(*))'/2y for all yE P_ 9?-, where
W+: (I+ -
QQ(*)) v2 p+y+
.1i_
and
W-: (I- -
KK(*))112p-y
_ 1+,.
are certain contractions. Moreover, 1) I1Kxll=11x11 11 W+(I+-QQ`*')x11 =11(1+-QQ(*')x11, 11=11y1111W-(I--KK(*)yll=1I(I--KKy11;
11Q
2) IIK11<1«II W+11<1,IIQ11<1«11 W-11<1; 3) (Y+, Y-) is a maximal dual pair if and only if W+:.ylt+ - QQ(*) C Ker W+ and
2
W : Ri- - KK(*) C Ker W- (Ritsner [4] Hint: See the definition of the symbol V ) in Exercise 7 to §8. In a Krein space Y= . + [ @]._N'- with infinite-dimensional JY construct a dual pair Y_) which is not a maximal dual pair and is such that + [+ 2'2'_ = .Y (Langer [7] ) Hint: See 10.10.
3
Choose the dual pair ('+, M'_) in Exercise 2 so that it has at least two different maximal extensions (Langer (71).
4
)Let 1=9o[+]1'1 [+] 92 and !Y1= (CA l)0 [+] (V 1)1 [+] (V l)2 be the decomposition of the subspaces 1 and 11 of a J-space W. Then 1o O+ (1'1)o, 1'1, (11)1 and 12 Q (11)2 are pairwise orthogonal and J-orthogonal subspaces completely invariant relative to the operator J:
[1o (@ (''1)o] = 1o (@ (1'1)0,
JC41= 11, etc.
Hint: Use the definition 10.3 of an (.,Y', .11- )-decomposition and the properties (7.1). 5
Under the conditions of Exercise 4
,0= [1o [+] 11 [+] 12] O [(1'1)o] + [(t1)1] + [(11)2], and the operator J is represented by the matrix J =
Aj 116;.1. Prove that: J41 = J14 is
an isometry, mapping (D 1)0 on to Do; 22 = J22 = 22'; J33 = J33, 0 < J33 < I; J63 = J36 =(I - J233) V36, where V36 is an isometry, mapping (11)2 on to 22; 55 = J55 = J55'; J66 = - V36 J33 V36; and J,j = 0 for the remaining i, j (Azizov). Hint: Use the result of Exercise 4 and the properties of the operator
J: J= J*= J-'.
6 7
Describe the maximal dual pairs (1+, M_) of subspaces from ./W' respectively. Hint: Use the result of Exercise 12 to §8. It is well-known that a basis [ ek) k -1 of a separable Hilbert space .)4 is said to be unconditional if it remains a basis under any rearrangement (i.e., renumbering) of its elements, and conditional in the opposite case. An almost normalized unconditional basis is called a Riesz basis (for an equivalent definition, see Chapter IV, §2.1).
78
8
1 The Geometry of Spaces with an Indefinite Metric Prove that a J-orthonormalized system .: = (ed k'= I in a separable Krein space N' is a Riesz basis if and only if the dual pair ('+, t-) generated by it is a maximal dual pair of uniformly definite subspaces (i.e., M'+ [+]. '_ =,,Y) ([VIII]). On the basis of the result of Exercise 7 give a complete description of all Jorthonormalized Riess bases in a separable Krein space.
Hint. By means of Lemma 8.4 and Theorem 8.17 describe all the canonical decompositions .yP = .1r+ [-+] -Y- of the Krein space and then use J-orthonormalized bases in .YY+ and .,Y- ([VIII]).
Remarks and bibliographical indications on Chapter I §1.1, §1.2. The systematic study of infinite-dimensional linear spaces with an Hermitian Q-metric was undertaken for the first time in a general formulation, it seems, by Nevanlinna [1]-[5]. Particular cases of such spaces (11', spaces) were considered earlier in the basic work of Pontryagin [1], M. Krein's articles
[3], [4], I. Iokvidov's articles [1], [2], and their joint papers [XIV], [XV]. General complex spaces . with a Q-metric were examined after Nevanlinna by
Pesonen [1], Louhivaara [1]-[3], Scheibe [1], Aronszajn [1], Ginzburg and Iokhvidov [VIII], M. Krein and Shmul'yan [1]-[4]. See the book [V] for a detailed exposition of the theory of these spaces, together with an extensive bibliography. We restrict ourselves here to only the necessary minimum of
information. Hilbert spaces with a real Q-form together with concrete applications are encountered in M. Krein's [3] (see [XV] ), and also M. Krein and Rutman [1], and in Hestenes [1]. §1.3. The concepts of a (Q,, Q2)-isometric isomorphism and of anti-space are well-known though formally introduced in [V]. § 1.4. For another proof of Proposition 1.9 see M. Krein and Shmul'yan [2]. The proof in the text is due to I. Iokhvidov. §1.5, §1.6. Here we follow [V] in our presentation. § 1.7. The maximality principle (Theorem 1.19) we encounter first in Scheibe [1], it seems. §1.8. Propositions 1.21 and 1.22 we find in Scheibe [1], and Proposition 1.23 in [V] but essentially also in [VIII].
§1.9. Propositions 1.24 and 1.25 belong to `mathematical folklore'; the proofs of them in the text belong to I. Iokhvidov and Azizov respectively. Proposition 1.26 is encountered in Pontryagin [1] and Lemma 1.27 in Ginzburg [3]. §1.10. The simplest case of finite-dimensional neutral skew-connected lineals and the theorem on the existence of Q-biorthogonal bases in them (cf. Lemma 1.31) we find in Pontryagin (1), and then in I. Iokhvidov [1] and in
[XIV]. There Proposition 1.30 (cf. Theorem 9.12) is proved and used systematically. The term 'skew-connected' lineals itself, which has become widely used, is due to M. Krein in [XIV]. Our exposition follows Bognar's [V], which, unfortunately, uses a different name `dual pairs', which now has in the literature (and in this book-see § 10) quite a different meaning. Corollary 1.32 and the notation SP, # Y2 have been borrowed from [V].
Remarks and bibliographical indications on Chapter 1
79
§1.11. The Example 1.33, due to Mackie, was published by Savage [1]; Theorem 1.34 is due to Ovchinnikov [1]. In this connection see also Markin [1] and Bognar [4]. The example of a non-degenerate lineal (1.17) which does not admit decomposition, due to M. L. Brodskiy, and also constructed on the difference of dimensions, but using different ideas as well, was published in [VIII]. §1.12. Lemma 1.35, which in such a general form is presented here for the first time, has a long prehistory, briefly traced in [XVI]. Our proof follows the
pattern in [V], where .5 =
and T= I.
§2.1-2.3. Essentially (although in an implicit form) Krein spaces were first
considered in 1954 by Nevanlinna [4] and then in 1956 by Pesonen [1), Louhivaara [1], and (independently) by I. lokhvidov and M. Krein (see [XIV], end of §13). Ginzburg [1], [2] began their systematic investigation (see below the remark on §2.4). In our exposition of the axiomatics and simplest properties of Krein spaces we follow Scheibe [1] and Bognar ([V]); the term `Krein space' itself is due to the latter. It should, however, be pointed out that in the book [V] (as also in the basic works of Pontryagin [1] and also [XIV], [XV] and some others-cf. [XVI]) the role of square brackets in the notation for the indefinite metric [x, y] is played by round brackets: (x, y). But we use round brackets to denote the (Hilbert) scalar product. This notation has been generally adopted in the last decades, especially in the Soviet literature. §2.4. The approach to Krein space considered here was first presented by Ginzburg [11, [2]. §3. All the results of this section, which here already become `folklore', can be found, essentially, in Ginzburg [1], [2]. To him also is due the extension of the term `J-metric', introduced earlier by Pontryagin for finite-dimensional spaces, to the general case. Example 3.11 is borrowed from [V]. §4.1, §4.2. The results of these paragraphs are due to Ginzburg [3] (see also [VIII]). The deficiency subspaces w = w' O P+ 2, _ ± (1(P±1) 1 were considered in such a general formulation (for arbitrary lineals /) by Ritsner [4]. In this connection see Exercises 7 and 8 to §8. §4.3. The results of this paragraph are due to I. Iokhvidov, in particular, in connection with Theorem 4.10 see I. lokhvidov [13]. Lemma 4.11 is wellknown and is applied widely (see, e.g., [XV], from which the construction of Example 4.12 has been borrowed; there it served quite a different purpose; in our context Example 4.12 is in a certain sense 'universal'-it is mentioned in §§5, 8, 9.
§4.4, §4.5. The results of these two paragraphs, moreover in a more general
form (in particular-for arbitrary spaces with a Hermitian Q-metric, and the `sufficient' part of Theorem 4.13-for spaces .T with a canonical decomposition .4'= .t + [-+] . -; see Exercise 6 to §3) were obtained by Scheibe (see also [VIII]). The proof given in the text of Proposition 4.19, which is suitable
for arbitrary spaces .: with a Q-metric, is due to I. Iokhvidov [VIII]. For a Krein space this Proposition follows from Ginzburg's results [3]. The formula (4.5) was published by Shmul'yan [4] with reference to an article of Bennevitz [ 11.
80
1 The Geometry of Spaces with an Indefinite Metric
§5.1-5.3. The concept of a uniformly definite lineal (in the original terminology a `regular definite lineal') was introduced in Ginzburg's investigation [4] (see also [VIII]). The term `uniformly definite' was introduced by M. Krein ( [XVII] ). Our presentation here follows the book [V]. Corollary 5.8 is due to Langer [9]. We remark that the definitions (5.1) and (5.2) of the intrinsic norm I I v, of a definite lineal 2 make sense in any space , with a Q-metric. They were so introduced (in another terminology in [VIII] (cf. [VI). §5.4. The classes h± were introduced (without special names) by Azizov [9] in connection with operators of the class .0, about which see below in §5 of Chapter III. More general classes of this same type were considered earlier by Langer [9]. §6.1. The example at the end of this paragraph is Ovchinnikov's [1]. For other examples, see Markin [ 1 ] .
§6.2, §6.3. A systematic investigation of the Gram operator Gv, was first undertaken in [VIII]. There too the term `Gram operator' was introduced and historical information about it was given. Theorem 6.4 in its complete form was established by Ginzburg [4] who indicated that he was following the ideas of Nevanlinna [4]. §6.4. The law of inertia (Theorem 6.7), although, it is true, not in such a
general formulation, we find already in Ginzburg
[3].
The rank of
indefiniteness xv, was formally introduced in [V], although essentially it was considered earlier by many authors (mainly in particular cases).
§6.5. Corollary 6.12 is contained in a more general proposition due to Ginzburg in [VIII]. §6.6. An extensive bibliography on W-spaces and G-spaces is contained in [III] and [IV]. Proposition 6.14 see in [III]. The first part of Proposition 6.13 was established essentially by Ginzburg [VIII] (see also Langer [21); the second part was pointed out by Azizov. The universality of Krein spaces was first established by Ginzburg [VIII]; the proof in the text was given by Azizov. Apart from the generalization of Krein spaces given in §6, this concept has been generalized in two other directions. The first of them relates to Banach spaces with a form Q(x, y) (I Q(x, y) I < (II x, 11 11 y II) continuous in them. The second considers normed spaces (in particular, Banach spaces) A " = A"+ [--] 4 --, where .4"± are arbitrary closed subspaces, P± are the
corresponding projectors, and an indefinite metric is given by the functional J"(x) = I I P+x I I " - I I P x II" (1 < v < co). For a short survey of these directions and their bibliography (up to 1970) see [IV]. §7.1. The material of this paragraph is traditional; our exposition follows M. Krein's lectures [XVII] and Ando [II]. §7.2. Lemma 7.7 is due to Ginzburg [4], but its proof in the text and also that of Proposition 7.6 was taken by us from [XVII]. §7.3, §7.4. The concept of projective completeness of a lineal was introduced independently by Scheibe [1] and in a particular case by I. Iokhvidov [21] (see [VIII]) to whom the term `projective completeness' is due. Lemma 7.11 can be found in Louhivaara [2], other results in §7.3 in Scheibe [1], and
Lemma 7.12 in Nevanlinna [4]. J-orthogonal projections, but only on
Remarks and bibliographical indications on Chapter 1
81
non-degenerate subspaces of the space II (see §9) were considered by Pontryagin [1], and after him by I. Iokhvidov [21] (see also [VIII]). But for spaces of a more general type (G-spaces and W-spaces) the development of the theory of projection was begun by Nevanlinna [4] and was later continued by
Louhivaara [2]. The final results were obtained by Ginzburg [4] (see also [VIII)).
§7.5. Regular subspaces (in a more general situation than ours) were introduced and investigated by Ginzburg [4] (see also [VIII] ). Theorem 7.16, summing up many investigations (cf. [V], [IV], and their bibliographies),
appears in such a general formulation here for the first time, apparently. Separate fragments of it have been proved by various authors in the course of almost 30 years. §7.6. Theorem 7.19 (there is' a particular case of it for [I spaces in [XIV]) is [1] (cf. also Theorem 2.1 in [VIII]. The proof given in the text is due to I. Iokhvidov.
essentially contained in a more general result of Scheibe's
§8.1, §8.2. The concept of an angular operator was introduced and investigated almost at the same time and independently in articles by Philips [1], [2] and by Ginzburg [3], [4]. The results in §8.1 and §8.2 are due to these authors. The term `angular operator' itself was first adopted in [VIII] at M. Krein's suggestion instead of the original name `angular coefficient' given by Ginzburg [3] (cf. Exercise 1 to §8). We borrowed the elegant version of the proof of Lemma 8.4 from [V]. §8.3, §8.4. The results in these two paragraphs have been widely applied after the publications of Philips and Ginzburg mentioned above; however, Proposition 8.9 with the remarks 8.8 and 8.9 connected with it, and Theorem 8.10 are published in connected form for the first time here, it seems. §8.5. In establishing Theorems 8.11 and 8.11' Ginzburg [3] did not find it necessary to have recourse to the definition (8.9) of the adjoint operator, since the angular operator K for' E W' (respectively Q for 2' E . ff-) was treated by him immediately as defined everywhere in .*', mapping .3W+ into W- (resp.
.W- into V+ ), and annihilating .,Y- (resp..+ ). However, later (cf. §§8.6, 8.7) such a treatment turns out to be inconvenient. Bognar [V] for the transition to the adjoint operator replaces, as we do, the angular operator K by the operator k = KP+ defined everywhere in W, and then considers k*. §8.6. Theorem 8.15, which is due to Azizov [3], was established by him for a
far more general case; it is referred to in [IV] (but again not in the most general formulation). Corollary 8.16 was proved earlier by M. Krein and Shmul'yan [3]. Similar results are obtained by Khatskevich [ 1 ] . §8.7, §8.8. The results in these paragraphs are due to Azizov; in particular Theorem 8.17 was published in an article by Azizov and Kondras [1]. The
classes d± appeared first in M. Krein's lectures [XVII]. They are systematically considered by Ritsner [4]. Later in Lemma 3.14 of Chapter III the reader will encounter a generalization of the classes 41. §9.1. Historically the spaces IIx (0 < x < oo) were the first infinite-
dimensional spaces with an indefinite metric to be studied by mathematicians. In the case x = 1 they were considered by Sobolev [1] (for the first reference to
1 The Geometry of Spaces with an Indefinite Metric
82
him see Pontryagin [1]), and for an arbitrary integer x > 0, on the example of the `model' separable space 12 (see Exercise 1 to §9), they were first studied by Pontryagin [1]. Here essentially all the results in §9.1 can be found, including Lemma 9.5. The proof of the latter in the text is due to I. Iokhvidov. Thus, although in this paragraph (see Definitions 9.1) the spaces II" are regarded as particular cases of Krein spaces, the latter themselves arose in the literature as generalizations of Pontryagin spaces (their axiomatics and the notation H,, itself are introduced in [XIV]; for more details about this, see [XVI]. §9.2. Theorem 9.6 was established by I. Iokhvidov (see [VIII] ); and the new proof given in the text, and Remark 9.8, are also his. Theorem 9.9 is essentially contained already in Pontryagin [ 1 ] .
§9.3. In Theorem 9.11 the implication b) - a), which is new compared with Theorem 9.6, is due to Ginzburg [4]. I. Iokhvidov [9] proved another (less effective) characterization of H,, spaces in the class of Krein spaces W: it is necessary and sufficient that any non-closed lineal 2' be complete relative to the intrinsic norm I I z- (cf. Exercises 11 and 12 to §9). §9.4. The formula (9.2) was estabished by I. Iokhvidov (see [XV], Theorem
4.1). The sharpening of it formulated in the second part of Theorem 9.12 is due to Bognar [V], as is the whole of the proof, as given in the text, of this theorem. §9.5. Proposition 9.13 is due to I. Iokhvidov. §9.6. W(')-spaces were considered in Azizov's article [7]. Proposition 9.15 and the first part of Corollary 9.16 are essentially contained already in [IV]; the second part of Corollary 9.16 are essentially contained already in [IV]; the second part of Corollary 9.16 and also Proposition 9.17 were noticed by Azizov.
Exercises to §9 Bognar's work [2] was the source for the results given in Exercises 9-12. §10.1, §10.2. Dual pains were introduced and studied by Phillips [3].
The first part of Theorem 10.2 is due to Phillips [3], the second part to Azizov. Other approaches to the description of extensions of dual pairs to maximal dual pairs are found in Phillips [3], Langer (Langer [7]; [V]) and Ritzner [3], [4]. Theorem 10.4 is due to Azizov. For certain applications of the theory of dual pairs not included in this monograph, see the literature cited in [V], Chapter V. §10.3. We find the first investigations of Q-orthonormalized systems in Nevanlinna's [3]; the most important of his results are reproduced in the survey [VIII]. J-orthonormalized systems were introduced and first investigated by Ginzburg in [VIII], and later-in great detail and in the more general scheme of G-orthonormalized systems (see §6.6)-by V. A. Shtraus [1], [2], to whom, in particular, Proposition 10.6 is due. All the rest of the exposition in §10.3
is due to
I.
Iokhvidov. We draw attention to the fact that our
terminology differs sharply from that of V. A. Shtraus [2] (more precisely, it is almost the exact opposite of his).
Remarks and bibliographical indications on Chapter 1
83
§ 10.4. Proposition 10.9 is due to V. A. Shtraus [2]; there too is published Proposition 10.10 (with an indication that is is due to Azizov). Remark 10.11 and the argument following it are published here by the authors for the first time.
Propositions 10.12 and 10.13 are due to V. A. Shtraus [2]. The proof of Proposition 10.12 given in the text is due to Azizov. We shall return to various bases in J-spaces in Chapter 4.
2 FUNDAMENTAL CLASSES OF OPERATORS IN SPACES WITH AN INDEFINITE METRIC
In this chapter the main classes of linear operators are defined and their simplest properties are described. As already mentioned, our main purpose is
to give an account of operators acting in Krein and Pontryagin spaces. However, in the process of studying the properties of these operators it turns out to be necessary to consider also more general spaces with an indefinite
metric, namely, W-spaces, G-spaces, and so on (see 1.§6). Most of the operators acting in these spaces have "definite" analogues. In the indefinite case we shall prefix the names with the letters W-, G-, J-, it-, etc., depending on whether the operators are acting in a W-space, a G-space, a Krein space (J-space), a Pontryagin space, etc. If an operator acts from one indefinite space into another, this will also be indicated by appropriate prefixes, whose meaning will be clear from the context. We mention, as will be clear from the size of the sections, that we devote most attention to the operators specific for an indefinite metric, namely, to plus-operators (§4). §1 1
The adjoint operator Tc Let .-W, and .JY2 be a G,-space and a G2-space respectively (see 1 §6.6) with
[, ]; _ (G,, )i (0 0 ap (G; ), i = 1, 2). The symbol T. W, -._A12 will in future mean that the linear operator T operates from a domain of definition VTC W, into a range of values RTC .IY2 the indefinite forms
84
§1 The adjoint operator T'
85
Definition 1.1: Let T:.H' -,W2, fir= .W'1. The operator T' defined on the lineal
(r = (y E .. 2 I there is a z E . i such that [Tx, y]2 = [x, z] 1 for all x E Jr) by the formula T`y = z is called the (G,, G2)-adjoint of the operator T. The connection of T` with the ordinary adjoint operator T* follows immediately from this definition: (1.1) T`= Gi -'T*G2. In particular, if Gi = I, is the identity operator in -Wi (i = 1, 2), then T' = T*. But if .W'i = 2 = . Wis ' a J-space, then
r = JT*J,
(1.2)
and therefore T' and T* are unitarily equivalent. From the relation (1.1) and, in particular, from (1.2) it follows that many properties of T* are preserved
for T'. 2 We give a geometrical interpretation of the concept of the adjoint operator T`, which will enable us later to prove easily a number of assertions. In doing this we shall for simplicity and brevity of exposition assume everywhere below in this section that Ai = Ye'2 = . is a J-space with the indefinite form [ , ] . 1.2: Every lineal 2' in the space of graphs x , f = ( (x, y) I x, y E k} is called a linear relation operating in .,Y, ,Or = with the domain of definition
Definition
v = (x I there is a y such that (x, y) E 2'), and the range of values
+Ir = (y there is an x such that (x, y) E 9), with
the kernal Ker 9= (x I (x, B> E 2'), and with the indefiniteness
Ind2'= (y I (0,Y) E 21.
In particular, if it follows from (0, y) E 2' that y = 0, then the relation T: x - y((x, y) E 2') is a linear operator, and 2' is its graph, which we shall denote but the symbol Fr = ((x, Tx> I X E 9r). In the space ,Yr we introduce the scalar product (1.3)
(<x1,y1), (x2,Y2))r=(xI,x2)+(YI,Y2).
It is well-known that I'r is a subspace in .Or with the scalar product (1.3) if and
only if T is a closed operator, i.e., if it follows from x
x0, Tx,, - yo that
x.E 9T and y,, = Txo. We introduce in -r the indefinite metric
[(xi,yi), <x2,Y2)lr=i([XI,Y21 - [Y1, X21)-
(1.4)
86
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
Hence, it follows that [
,
}r = (Jr-, )r, where Jr(x, y) _ < - iJy, iJx), and
therefore Jr = Jr = Jr'. Consequently the space fir admits the canonical decomposition
Wr=. °i
0+
r,
(1.5)
where .Ye
r = Ptr .YPr,
Prr = z (Ir ± Jr ),
and therefore
.Wr =((x,±iJx)IxE.
(1.6)
}.
Let T be a linear operator acting in .', and let rT be its graph. Then
Pr rT= (((I + iJT)x, ± iJ(I + iJT)x) I xE CAT}.
(1.7)
In particular, if r r is a non-negative lineal, then its angular operator KT can be written symbolically in the form of a pair
KT= < V, - JVJ),
(1.8)
V= (I+ UT) (I - iJT)-';
(1.9)
where
here it is borne in mind that the operator I - UT is necessarily invertible and this means that
KT( x, y) _ (Vxi - JVJy)
(1.10)
for (x, y> E 9K,1:1 From Definition 1.1 of the operator T' and from the Definition 1.1.11 it follows immediately that
1.3. If T. Y(
.Yf is a linear operator with t T =
, then r ri = F P.
Corollary 1.4: In order that a vector xo should belong to VT n Vr and that Txo = Tcxo it is necessary and sufficient that the vector (xo, Txo> be isotropic in I 'T.
The condition imposed on xo is equivalent to the fact that <xo, Txo) E F T n r,- = FT n rV1, i.e., (xo, Txo) is isotropic in rr by the Definition 1.1.12.
3
From Proposition 1.3 it also follows that:
1.5
The operator T` is closed.
1.6
If for a densely defined operator T there is a densely defined inverse
operator, then (T-')`= (T`)-'.
§1 The adjoint operator T`
87
1.7 If an opeatror T2 is an extension of an operator Ti (Ti C T2) and if VT, = .h , then Tz C Tic. 1.8
In order that the operator T" _ (T`)` should exist it is necessary and
sufficient that VT =,,Y and that the linear operator T admit closure; in addition,
T` = T' T" = T, and, in particular, t = T. Proposition 1.5 follows from the fact that the J,-orthogonal complement is a subspace (see I, Proposition 7.1). Proposition 1.6 follows from Proposition 1.3 if we note that F,-' _ ((Tx, x) I x E 1'r). To prove 1.7 and 1.8 it is necessary to use 1 Formula (1.8) and 1 Proposition 7.4 respectively. It follows from (1.2) that in a Krein space the operator T` is bounded only when the operator T is bounded. We note also a number of other properties of T`.
and CtT, n CIr, =.Ye. Then (T1 + T2)` J T i + T'2, and 1.9 Let Ti, T2: ,YP if at least one of the operators T, or T2 is continuous and defined everywhere,
then (T1+ T2)c = Ti+T`2. 1.10
Let T1iT2:,Ye -', and 1'r2=
r2r,=ye. Then (T2T1)`DTiTz,
and if either T2 is bounded, or if the operator Ti', defined everywhere and bounded, exists, then (T2T1)` = TiTz. We verify only the last assertion, the others are verified analogously. Thus suppose that Ti- 1 is a bounded operator and *r, = W. Since we are regarding
the inclusion TiTz C (T2 Ti )` as proved, it remains only to verify that (T2T1)` C TiTz. Let xo E d(r,r,). Then [T2T1x, xo] = [x, (T2T1)`xo] _ [Ti'Tix, (T2T1)`xol
= [Tlx,(Ti')`(T2T1)`xo] for all xE 'T27-,. Since 91 r, =, , so 9r2 = Tl'r2r and therefore xE Sri and Proposition 1.6, we Using obtain Tzxo= (T,-')`(T2T1)`xo TiTzxo = (T2Ti)`xo.
From the Definition 1.1 of a J-adjoint operator T` it follows
4
immediately that: 1.11
If a subspace if is invariant relative to the operator
T,
i.e., if
n CAT= 9 and T( J r n Y, then
Tc(Vr, n
1J) c YILI.
Since _r' for a closed operator we always have Ker T = 4? T. and the relations
and iRr=fir are eqquivalent (see, e.g. [XIX] ), it follows from the RT' =. equations fir' = J d l r ' and Irl' = (J: T')[11 = J. J,ylrl] = 3 J (= Ker T), in which 1, Propositions 7.1 and 7.3 have been used, that: 1
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
88
1.12
If T is a closed operator and 5l r = M, then always Ker T = R7"-", but
Rr` _ (Ker T)I1l if and only if Rr=fir. Let The a linear operator and X one of its eigenvalues. We recall that the set of all vectors x in the space, for which there exists a natural number p(x) such that (T - XI )P x = 0, forms a lineal of `root vectors corresponding to the point X' (briefly, a root lineal), and it is denoted by the symbol . f,(T). The following theorem characterizes the disposition, in the field ,3'f, of the invariant subspaces of the operators T and T`, and also their root lineals.
Theorem 1.13: Let T: °- 4e, let GOT= W, '( C fir) be an invariant subspace of the operator T, let V( C mss) be on invariant subspace of the operator T`, and let a(T 121) fl a*(T` I ,N) = 0. Then 21 [L] .jF. This implies, in particular, that 21 (T) is J-orthogonal to 21 (T`) when X ;4µ. Let X E 2', y E IV. We bring into consideration the following functions of the complex variable X: g.r.y(X)= [(TIC-XI) 'x,Y] and gXy(X)= [x,(T`I.N-AI)-'y] The first of these functions is defined and holomorphic (because the resolvent is holomorphic) on the set p (T 121) of regular points of the operator T 12', and the second on the set p*(T` I . IV). Let I X I /> max (II T 19? IIII , II T`.IY II 1. Then X 9)
gXy(X) _ [(TI Y- >,I)-'x, y] = l - Z 1
-
TI - X,YJ 1
X "70
L\-XTIY)"x, y].
Similarly we obtain that
gXy(X)=-1 Z IX,(-1 T`IA) nyl Since
RX,Y] =[x,I we have gry(X) = g,'(y(X) when I X I > max(II TI 2' I , I I T` I a ' ) , i.e. each of these functions admits analytic continuation gx,y(X) on to p(T 12) U p*(T` I A'). From the condition a(T 12') fl a *(T` I ,i") = 0 we conclude that p(T 12') U p*(Tc I A") = C and the function gg,y(X) is bounded on 1
1
I
C. Therefore, by the well-known Liouville theorem, g,,y(X) = const. In combination with the fact that g,,,y(X) 0 as I X I oo, we obtain that gx,y (X) = 0. In particular, for X E p (T 12') and arbitrary x E 2' and y E .4' we 0, i.e. 2' [L] ./V. have [x, y] = &l v We now prove that Yx (7-) [L] 7, when X # µ. Since the form [ , ] is continuous, it suffices to verify that -T,(T) [L]
LetxE2),(T), yE2 (T`). We put 21= Lin((T - XI)"x(o,
. t"=LinI(T`-td)°Y]o
§1 The adjoint uperator T`
89
These are finite-dimensional invariant subspaces of the operators T and T respectively, and a(T I 97) = (X), and a(T` I.i") = (µ). Since X * µ by hypothesis, so a(T I 9') fl a* (T` I A) = 0, and from what has been proved above [x, y] = 0, i.e. 9,,(T) [l] .7,,(T`) when X ;4 A. Before going on to the next proposition we recall some definitions and notation. Definition 1.14: Let T be a closed linear operator in an arbitrary Hilbert space. A point Xo is called a normal point of T (Xo E p(T)) if it is either one of its regular points or if it is a normal eigenvalue (Xo E op(T)). The latter means that there is a decomposition of the space = 97 + .if" into subspaces 91 and./{"such that 9"C 1r, dim 97< oo, T: 97 97, T:.,t" a(T I [F)_ {Xo) and XoEp(T I A").
Hence, in particular, it follows that: 1.15 Normal eigenvalues are isolated points of the spectrum, and therefore in every open set they make up no more than a countable subset.
The symbols ap,, (T) and ap,2 (T) denote those X E ap (T) for which
r-u # . and.;tr-u = .,Y respectively. It is clear that ap,, (T) fl ap,2(T) = 0 and ap(T) = ap,, U ap,2(T). Therem 1.16: Let .Yt' be a Krein space and T: r C T = M. Then
,Y' a closed operator with
a) XEp(T)aXEp(T`); b) XEa,(T)aXEap,2(T`); c) XEap,,(T) a XEap,l(T`);
d) XEap(T)aXEap(T`); and, as a consequence of a)-d), e) X E a(T) 0 X E a(T`); f) XE ac(T) 0 X E ac(T`); The proof of these assertions follows immediately from formula (1.2) and the fact that analogous assertions hold for the spectra of the operators T and T*. As an example, we examine the proof of the implication d). By 1.8 it suffices to verify that [X]) X E ap(T) * X E ap(T`). Since (see X E ap(T) - X E ap(T*), so it then follows by Definition 1.14 that there is a decomposition M _ 97 + A" such that Se C 1 r*, dim .7 < oo, T*: 9 9),
T*: v- V, a(T* 19') _ (X) and I E p(T * 14"). Hence, we obtain without that de = J9' + J,4, r: J'4' - JA Tr J9' C J2', a(Tc I PT) = a(T* 9')( = {N)), p(T` I J.N) = p(T* I A), and therefore difficulty
X E ap(T`).
Now let T be a linear operator acting in a W-space YP, and let YY° be an isotropic subspace of YY, .YP° C VT and TWO C .W0. We form the factor5
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
90
space .
= .W/.° (see 1, §1.8) and bring into consideration the operator T:
/'T= {zI xE /r),
Tx= Tx
(1.11)
It is not difficult to verify (see, e.g., [XIX]) that the operator T is properly defined. Moreover, closedness (respectively boundedness) of T in the norm II x II generated by the scalar product (x, y) implies the closedness (respectively
boundedness) of t in the norm
I x II " = inf { II x II I x E z) which, also is generated by the scalar product (x, y) = (x, y) where x E z, y E y and x, y are implies t = .Yf. orthogonal to 9'o; and r =
Let k be a Krein space, T a linear operator, 9T = .; let 910 be a neutral subspace contained in c/T n cT° and let T. ° C Y°, T`20 C 2°. We form the = F'/2° and consider in it the Krein space (see 1, Corollary 5.8)
operators t and T` generated by the operators T and T` respectively in accordance with formula (1.11). The following theorem holds: Theorem 1.17:
(r
The density of cr in .l° was noted earlier. Therefore the operator T` is defined (see Definition 1.1). Since Yo C c/ r, so z E cr if and only if x E ('r. Hence, if x E (r and y E C17, then (1.12) [z, T`y]A= [TX,YIA= [Tx,y] _ [x, T`y] = [X, f'9] A, i.e, T`y = T` y, and therefore it remains to prove that V j-' D VT. Indeed, if y E VF, , then, reading the system of equalities (1.12) from right to left, we obtain that y E (/ r.
Exercises and problems 1
Verify that Definition 1.1 is correct if the condition c/T = , in it is replaced by the condition of `G1-denseness' of VT: DT 1 = (0), and prove that a) T` is a closed
operator; b) if it is G2-dense and Ker T= (0), then (T- ')` = (T c)-' [III]. lTi+T. = 9T. (17,r. = (r.) and r/j- =.Y2, the relation (T, + T2)` = T, + Tz ((T,T2)` = T 2 holds ([III]).
2
Prove that when
3
Prove that if, for any T: _Wj -' W2, the G1-denseness of VT implies the G2denseness of c r, then 0 E p(Gi) ([III]).
4
Investigate which of the implications in Theorem 1.16 continue to hold in the case of G-spaces with 0 O p(G) ([III]).
5
Let T be a linear operator bounded in II, = H+ [ + ] I- F. . Prove that, if the space n, is not separable, then there is a separable subspace II,( C 7r.), invariant relative to T and T`, such that Ilk = II; [ + ] Ilo, where Ho is a Hilbert space relative to the
scalar product- [x, y] and that the operator T is represented relative to the indicated decomposition by the matrix
T=
II
T'
0
0
To
I
([XVII])
§1 The adjoint operator T`
91
Hint: Fix in II+ any basis el, e2...., e,,, and choose as II; the closed linear envelope of all vectors of the form T" I (T`)"T"3 ...T 1(T `)n2,ek, k= 1, 2, ..., x; 0
np < oo, p = 1, 2,... .
Let H ; be J;-spaces (i = 1 , 2), T:.Yi - Y2. We define the space W3 = 0 1 + (D IYZ and the operators Pi + PZ T: .)Yi -W3, Pi + P2 TPi : J e - 3,
6
Pi TP; : r'1 -,N2. Show that the following conditions are equivalent:
a)KerPzTP; I.e =10);
b) Ker (Pt + Pz TPi) = (6); c) Ker (Pi + Pi T) = (0) (cf. I. Iokhvidov [14], [18]). Verify that under the conditions of Exercise 6 the following requirements are
7
equivalent: a) .Pz TPi = 2 ; b) RP, + P: TPi = N3;
c) RP,+ PZ T = 1Y3 (cf. I. Iokhvidov [14], [18] ).
Independently of Theorem 1.13 verify that -149)(T) [I] when X ; µ (cf. Pontryagin [1]). Hint: Consider the minimal non-negative integers p and q such that (T - XI)px = 0 and (T`-µl)Qy=0 if xEYx(T) and and apply induction over p + q.
8
Dissipative operators
§2
Definition 2.1: A linear operator A with an arbitrary domain of
1
definition Q)A, operating in a W-space .9, is said to be W-dissipative if Im [Ax, x] > 0 for all x E 9A, and to be maximal W-dissipative if it is W-dissipative and coincides with any W-dissipative extension of it.
In particular, when W= I this definition brings us back to the (usual) dissipative and maximal dissipative operators in ,Y (see [XI], [XVIII] ).
From Zorn's lemma easily follows the Proposition 2.2
Every W-dissipative operator can be extended into a maximal one.
Now let A be a J-dissipative operator. From Definition 2.1 we have 2.3 An operator A is J-dissipative if and only if each of the operators JA and AJ is J-dissipative. Moreover, A is a maximal J-dissipative operator if and only if each of the operators JA and AJ is a maximal J-dissipative operator. The proof of the second part of Proposition 2.3 uses the fact that an operator T is an extension of an operator S if and only if the operators JT and TJ are extensions of the operators JS and SJ respectively.
It is easy to see that an operator A is J-dissipative if and only if its graph rA in .0 X ,,Y in the metric (1.4) is non-negative. 2.4 Let A be a densely defined J-dissipative operator, and let the lineal Y be a
non-negative extension of its graph in . e x W. Then ' is the graph of a J-dissipative operator with is an extension of the operator A.
92
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
In the proof here we need only the proposition that Y is the graph of some operator, i.e., that (0, y) E . implies y = 0. Since Y is a non-negative subspace in [(0, y), (0, y) ] r = 0, <0, y) is an isotropic vector in 2' and therefore it is Jr-orthogonal to FA, i.e., [y, x] = 0 for all X E A. Since 2A = . it follows that
y=0. Corollary 2.5: A densely defined J-disspative operator A admits a closure which is J-dissipative, and therefore, if A is a maximal J-dissipative operator W, then A is closed. and
This follows from 1 Proposition 3.6.
Theorem 2.6: A closed J-disspative operator A admits maximal closed J-dissipative extensions. Moreover the latter are densely defined and their graphs are maximal non-negative subspaces in ,Y x .. Since a non-negative subspace in a Krien space can be extended into a maximal subspace by adding a uniformly positive subspace orthogonal to it (relative to the indefinite metric) (see I (8.8)), we shall prove that such an extension f of the graph IPA in W x W is the graph of a J-dissipative closed extension A of the operator A. For, in the opposite case there would be a vector (9, yo) in f with yo * 0. Since such a vector is neutral by the form (1.4), it is isotropic in F, but by the construction of f its isotropic part coincides with the isotropic part of IPA; but rA is the graph of the operator A and therefore yo = 0-a contradiction. Thus, every closed J-dissipative operator A admits extension into a closed J-dissipative operator A. whose graph is a maximal non-negative subspace in
.W x M. Hence it follows that A is a maximal J-dissipative operator. It remains to verify that it is densely defined. Let yo be J-orthogonal toA. Then the neutral vector (B, yo) is Jr-orthogonal to PA. Therefore yo = 0. We now dwell on other criteria for dissipative operators to be maximal.
2.7 A closed J-dissipative operator A is maximal if and only if 1 A = .
and
(- A`) is J-dissipative, and therefore the operators A and (- A`) can be maximal J-dissipative only simultaneously.
Taking Proposition 1.3 and Theorem 2.6 into account, this assertion coincides with Theorems 8.11 and 8.11' in Chapter 1. Before formulating the next result we recall a definition. A point X E C is called a point of regular type of a linear operator T: e yt' if the operator (T- XI)-' exists and is continuous. The set of all regular points of an operator T is called its field of regularity. We point out that here neither the operator T nor the operator (T- XI)-' is assumed to be defined on the whole of W. Lemma 2.8:
Let T be a closed dissipative operator in a Hilbert space .,Y.
§1 The adjoint operator T`
93
Then the open lower half-plane C- belongs to the set of points of regular type
of the operator T and II (T- XI)-' II <, I/ I Im X I (X E C-). Moreover, T is a maximal dissipative operator if and only if c- fl p(T) * 0 or, equivalently, C- C p(T). If Im X < 0, then it can be verified directly that, since T is dissipative ((T- XI)x, (T- XI)x) '> Im2 X(x,x) and therefore I (T - a1)-1 II I/ I Im X I . Consequently X is a point of regular type of the operator T. Therefore, if C- fl p(T) ;4 0, then T is a maximal dissipative operator. In the opposite case X E c- fl p(T) would be an eigenvalue of a dissipative extension T D T. To complete the proof of the lemma we verify that if T is a maximal
dissipative operator, then C- C p(T). For, suppose xo E 91¢_xf for some X E C. If xo # B, then the vector (xo, txo) is positive by the form (1.4) and is Jr-orthgonal to the maximal (by virtue of Theorem 2.6) non-negative subspace Pr, which is impossible, and therefore xo = Oi i.e., , ?T-1,i = '. Since (T- XI)-' is bounded and T is a closed operator, so X E p(T), i.e.,
C- C p(T). We now suppose that a maximal uniformly positive subspace Y is contained in the domain of definition of the operator T. By Lemma 9.5 of Chapter 1 this supposition is always fulfilled if , is a Pontryagin space H. Taking account of §7.6 of Chapter 1 we shall, without loss of generality, suppose that ' = M'. Then, relative to the canonical decomposition , = Y" + ,-, the operator T is expressible in matrix form (2.1)
T=II T%iII?1=1
where T,1=P+TP+I,
+, T12=P+TP-IJr-,
T21=P-TP+IW+,
T22 = P- TP- I 'W_.
For example, J= 11
J1iII'.i=1, .Iii=1+, J12=0, J21=0, J22=-I-.
Theorem 2.9: Let A be a closed J-dissipative operator, following conditions are equivalent: a) b) c) d)
A is a maximal J-dissipative operator; - A 22 is a maximal dissipative operator in)
+ C VA. Then the
;
[XIIm X>2IIAP+II flp(A)Pe 0; ( X I Im k> 2IIAP+II) Cp(A)
If these conditions are satisfied, then II (A - XI)-' II = O(1 /Im X) (Im X - co ).
Let A be a maximal J-dissipative operator. By Proposition 2.3, JA is a maximal dissipative operator. It is easily verified that, since JA is dissipative, (JA)22 = -A22 is also dissipative. We prove that it is maximal, by checking
94
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
that C- n p( - A22) # 0 (see Lemma 2.8), or, what comes to the same thing, that C+ n p(A22) ;6 0. Since
when JAP-+XI=(I- JAP+ (JA + XI)-')(JA + XI) XEC+, so, when Im X > II AP+ 11 we have (see Lemma 2.8) - X E p (JAP- ), and since
(-A22+ XI-)-' = P-(JAP- + XI)-'P- I, -, so XEp(A22) Now let (- A22) be a maximal dissipative operator in
-. Since W+ C VA,
every non-trivial J-dissipative extension of the operator A would imply a non-trivial dissipative extension of the operator (-A22), and therefore A is a maximal J-dissipative operator and AJ is a maximal dissipative operator. Therefore, from the equality
A - XI= -(I-2AP+(AJ+ XI)-')(AJ+ X1) (which is proper because J9A = 9A), we obtain, taking account of Lemma 2.8, that, if Im X>2 1 1 AP+ II, then X E p(A), and moreover
II (A - XI)' II=0(1/Im X)when Im X, oo. Thus we have verified the following implications: a) a b) - d). The implication d) s c) is trivial. We verify that c) - a). If Im X > 2 II AP+ II and X E p (A ), then A cannot admit closed J-dissipative extensions A l for otherwise we would have X E aa(A), which contradicts the implication a) - d), which has already been proved.
Definition 2.10: Let T: Ye, dY2 and S:,'1 --' ,W1 be two linear operators. The operator T is said to be S-bounded (or S-continuous) if even for one point Xo E p(S) (but then, as follows from Hilbert's identity for the 2
resolvent, also for all points) the operator T(S- XoI)-' is bounded and defined everywhere. If, in addition, T(S - X01)-' is a completely continuous operator (T(S - XoI)-' E ,9 ), then the operator T is said to be S-completely continuous. In particular, as is easily seen, bounded (respectively, completely continuous) operators A with VA =1 are S-bounded (respectively, S-completely continuous) for any operator S with p(S) ;d 0. _
A closed operator T: 1-1,'2 is called a cF-operator if 1T=?T, dim Ker T < oo and dim Rj < oo; and the number ind T = dim .Wj - dim Ker T is called its index. In particular, if ind T = 0, then such an operator is called a Fo-operator. Let A: ,Y - M, Xo E C, and let T = A - X01 be a (Do-operator; then the point Xo will be called a cFo-point of the operator A. We observe that, since .3111 = Ker T* when fT= W1, the definition of a cF-operator can in this case be given formally in a different way: namely, by replacing R# in the definition by Ker T*. We shall use both these definitions in future.
We note also that if A: W .Jf and if Xo is a normal point of the operator A, then Ao is a cFo-point of this operator. In particular, the 'Fo-points of any completely continuous operator fill the set C \ (0]. We state without proof a theorem made up of several results form the survey [X] which are applicable to our present interests.
§2 Dissipative operators
95
Theorem 2.11: Let TI: .W, -.,Y2 and T2:.W2 -.03 be 4'-operators, and J3 is a (D-operator and ind T2T1 = ind T1 VT2 = . 2. Then T2T1: + ind T2. If also T3: W1 -02 is a completely continuous operator, then '2 is a 4)-operator and ind(T1 + T3) = ind T1. T, + T3: 1
1
Let T:
.
.i( be a 4)o-operator. If T is an S-completely continuous
operator, then T + S is also a 4)o-operator. The set of 4)o-points of the operator T is open. If ( is a connected component of this set and if Sl n p (T) ;d 0, then
0 C p(T). Corollary 2.12: If A is a closed maximal J-dissipative operator, .-W+ C 9n, then A 12 is on A22-continuous operator. If also A 11 - A 11 E 9, and A12 is an A22-continuous operator, then C+ C p(A).
Let X E p(A) fl
C+.
Then AP - - XI = [I - AP+ (A - XI)-'] (A - XI),
and since by Theorem 2.9 11 (A - XI) - ' 11 = 0(1/Im X) when Im X - oD, so C+ fl p (AP-) 0. This implies, by Lemma 2.8, the inclusion C+ C p (AP- ). we have, by Definition Since [(AP- - XI)_ 1112 = (l/X)A12(A22 2.10, that A12 is an A22-continuous operator. Let X0 E C+. Since All - XoI+ _ (2' (A + A) - Xol+) + z (A 11 - A *1) and Xo E p(z (A + A*1)) but A - A 11 E .9 , so, by Theorem 8.11, Xo is a 4)o-point
of the operator All. Moreover,
since All is
a bounded operator,
C+ fl p (A 1,) ;e 0, and therefore C+ S p (A 11). We now represent the operator A - X01 by the expression
A - Xol+ A21
0
I-
0 A12(A22-XOI ) - ' 0
01
x
I+
0
0
A22 - Xol
The first term in the round brackets and the second factor are 4)o-operators,
and the second term in the round brackets is a completely continuous operator. It follows from Theorem 2.11 that A - XoI is a 4)o-operator, and from Theorem 2.9 it follows that C+ fl p(A) # 0, and therefore (see Theorem 2.11) C+ C p(A). Corollary 2.13: A closed a-dissipative operator A in fl, =11+ G) 171- is maximal if and only if C+ fl p (A) # 0, and in that case C+ C p'(A ). O
Let A be a closed maximal a-dissipative operator. It follows from
Proposition 2.7 that 'A = I1; and therefore (see 1, Lemma 9.5) we can assume without loss of generality that 11+ C 9A. By Corollary 2.12 we have that A12 is an A22-continuous operator. Since the operators All and A 12 (A22 - XoI-) - 1 (Xo C p (A22)) are finite-dimensional and continuous, they are completely continuous, and therefore C+ E p(A). Conversely, let Xo E C+ fl p(A). Then from the equality
AP- - XoI =(I - AP+ (A - X01)-')(A - X0I )
96
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
and the facts that AP' (A - XoI)-' is finite-dimensional and that c- fl op(AP-) = 0 it follows that Xo E p(AP- ), and therefore Xo E p(A22). It remains to apply Lemma 2.8 and Theorem 2.9, from which it follows that A is a maximal .7r-dissipative operator.
In this paragraph our main purpose is to study the structure of root lineals of dissipative operators in indefinite spaces. With every densely defined linear operator T. W Xe we associate the pair of operators
3
TR ='(T+ T`) and T, = Zi (T - T`) which we shall call the J-real part and the J-imaginary part of this operator respectively. We note that cTR = cT, _ gT n CAT', and if X E t - fl VT', then Im [Tx, x]
Tix, x]
(2.2)
Consequently the following holds: 2.14
If 2A C 9A°, then the operator A is J-dissipative if and only if
[Aix, x] > 0 (XE 9A). In particular, this is true if A is a continuous operator defined everywhere in W.
Theorem 2.15: Let A be a J-dissipative operator, 1A = e. Then the relations Im [Axo, xo] = 0 and xo E Ker A, (i.e., Axo = A`xo) are equivalent. If xo E Ker A,, then Im [Axo, xo] = 0 by 12.2).
Now suppose Im [Axo, xo] = 0. Then the vector <xo, Axo) of the nonnegative subspace rA is neutral relative to the form (1.4), and therefore it is isotropic in rA (see 1, Proposition 1.17). It only remains to use Corollary 1.4. Theorem 2.15 gives a basis for introducing the following notation in the case of an arbitrary W-dissipative operator A: Ker A, = (x I Im [Ax, x] = 01,
(2.3)
although the symbol A, has no meaning in such a general case. From Theorem 2.15 we obtain Corollary 2.16: Let A be a J-dissipative operator, VA = .. Then Ker (A - XI) C Ker(A ` - xI) fl Ker A, for all X _ X .
If xE Ker(A - XI), then Im [Ax, x] = Im X [x, x] = 0, and therefore our assertion follows from Theorem 2.15 and the definition of the operator A,.
§2 Dissipative operators
97
Corollary 2.17: If A is a maximal closed J-dissipative operator, then Ker (A - V) = Ker(A c - V) when X = X and therefore the residual spectrum of the operator A contains no real points.
By Proposition 2.7, A and (-Ac) are maximal J-dissipative operators simultaneously. Since A" = A (see Proposition 1.8), it follows from Corollary
2.16 that the kernals of (A - XI) and (A` - XI) coincide: Ker(A - X) = Ker(A ` - XI) when X = X, and from Theorem 1.16 b) that the operator A has no real residual spectrum. We observe that the the set introduced in formula (2.3) coincides, as may easily be seen, with the set (x E 9A I [Ax, y] = [x, A y] for all y E ftA I Therefore, carrying out an argument analogous to that used in Exercise 8 to § 1 we obtain
Corollary2.18: Let A be a W-dissipative operator, let 2' C 14(A) fl Ker AI, and AY C Y. Then the lineal 2' is W-orthogonal to all l,,(A) with µ * X. Lemma 2.19:
Let W° be an isotropic subspace of a W-space L, let A be a
W-dissipative operator, A = Je, and .° C 9A. Then AX° C .'°. It is easy to see that r° C Ker AI. Therefore [Ax, y] _ [x, Ay] for any x E M° and Y E 9o. Since A = M, AXE .W°.
4
Before proceeding with the exposition of the properties of dissipative
operators in spaces with an indefinite metric we recall the definition of a Riesz projector and some of its properties.
Let T be a linear operator in an arbitrary Hilbert space .', let a be its bounded spectral set (with the meaning given in [VI]) i.e., the isolated part of the spectrum (or, in particular, the whole spectrum) of the operator T, and let F. be a Jordan closed rectifiable contour lying in p(T) and containing a, and suppose a(T)\a lies outside this contour. Then the operator
(T - XI)- dX, (2.4) r, where the integral is understood to be the strong limit of the integral sums, is called a Riesz projector. We enumerate the basic properties of the integral (22). Po
Theorem 2.20: a)
1
2ai
The following assertions hold:
the operator P,, does not depend on the choice of the contour F. isolating the set a, and it is a projector;
b)
PQ.I C CA T;
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
98
c)
the subspaces P
and (I - PQ).
are invariant relative to the operator
T, and also a(T I Pa') = a, a(T I (I - PQ)) = a(T)\a. In particular, if a = (Xo) is a normal ergenvalue of the operator T, then _ Iao(T); d) if T is a bounded operator and a = a(T), then P. = I; e) if a, and a2 are bounded spectral sets of the operator T and a, fl a2 = 0,
P
then PQ, U a = PQ, + PQ, and P., P, = PQZ PQ, = 0;
f) if a is a bounded spectral set of the operator T, and ? is an invariant subspace of this operator, and a (T I 2) C a, then 2' C P0.. Proofs of these assertions can be found, eg, in [VI]. Theorem 2.21: Let ' be a W-space, A a closed W-dissipative operator, a( C C+ (C- )) be a bounded spectral set of A, and let P. be the corresponding Riesz projector. Then P0W is a non-negative (non-positive) subspace which is invariant relative to A. The proof will be carried out for the case a C C+, since the case a C C- is
treated exactly analogously. Without loss of generality we can assume, by is a virtue of assertions b) and c) in Theorem 2.20, that A: P - P bounded operator and a(A) = a. Then from assertion d) of the same theorem we have P,, = I. Therefore, [x, x] = [Pox, x] and it suffices for us to verify the inequality Re[Pax, x] > 0. To do this we choose as F. the contour consisting of a segment [ - a, a] of the real axis and a semicircle (ae")'=o with a sufficiently large a > 0. Then
P
1
27ri
f -aa (A-aI)-'da+2a
Jo
(I- ea
Since P. does not depend on the contour PQ separating a, the
limit
lima -. - Pa = PQ( = I) exists. Since Um
1
('x (J_iA)- ' d,p=ZI,
aco2lr J o
a
so there exists also ra
lim
aim
Q (A - aI) ' da=
27r
(A-cI)
da=z I.
Therefore Re[PQx, x]
2Rel r L[27r
Jm
(A-aI) dax,x] = - I f -- in[(A-aI)-'x,x] da>, 0. 7r
The last inequality follows because A is a W-dissipative operator.
§2 Dissipative operators
99
Corollary 2.22: Let A C C+ (C-) be a certain set of eigenvalues of a W-dissipative operator A. Then C Lin (2 (A) l x c A is a non-negative (nonpositive) subspace. As in the proof of Theorem 2.21 we assume for definiteness that A C
C+,
and by virtue of 13.6 it suffices to verify that Lin (2'), (A) l x E A is a non-negative lineal. Let x E Lin (2'>,(A) la E A. Then there are X,, X2, ..., X E A, and positive
integers P,, P2, ..., and vectors x,, x2, . . ., x such that x = E°=, xi, (A - X,I)°ix, = 0. The subspace 2' = Lin(xi, (A - X,)x,, ..., (A - A,I)°'-'x;l; is finite-dimensional, is contained in c/A and is invariant relative to the operator A, moreover, a(A 19) = (X,l i C C+. By Theorem 2.21 the subspace 2' is non-negative, and therefore the vector x E 2' is also non-negative.
Let A be a maximal ir-dissipative operator in H, with c'A = II,,. The a(A) (1 C+ consists of not more than x (taking algebriac
Corollary 2.23:
multiplicity into account) normal eigenvalues.
The proof of this assertion follows directly from a comparison of Corollaries 2.13 and 2.22 with Proposition 9.2 in Chapter I. ±
±
Theorem 2.24: If A is a closed J-dissipative operator, if a (C C ) are its bounded spectral sets, and if FQ= C C ± and r, = r a*- (= (X I X E r,,- J ),then the subspace PQ-uQ-.W is non-degenerate. Let W° be an isotropic subspace of the subspace PQ - uo -.iC. By Theorem 2.20 Po - uQ ' C cA, and even more so ,0° C VA, and therefore (see Lemma 2.19) A.° C e°. Since W° C Ker AI and, as is easy to see, so (A - XI)-' I. ° = [(A - XI)-']` jW° when Fa-uC- C p(A I.W°), X E ro-uQ-.
(For, [Ax, y] = [x, Ay] for any x E MO and y E /A. Therefore
[ (A - V) x, y] _ [x, (A - XI )y] Now let
x=(A- XI)-'z (zE Y(°) and y=(A-XI)-'w (wE.f). Then
[z, (A- XI)-'w] = [(A - XI)-'z, w]. Since this is true for any w E, we have [(A - XI)-'] `z = (A - XI)-'z for any z E .°.) Since
Pa'Ua- _ -
1
27r:
r,-,,-
[(A - XI)-']` dX,
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
100 so
Po-ua-I.-,Y°=P,-uo-I'°=II-W°.
[x, xo] _ [x, PQ-ua-xo] = [Pa-ua-x, xo] = 0 for any xE J' xo E Je°, i.e., J 0 is isotropic in the whole of Y. Therefore 0 = (0). Hence
and
Let A be a J-dissipative operator, and let the non-real X and X be its normal points. Then 2' = Lin 191),(A), 21,(A)) is a projectionally
Corollary 2.25:
complete subspace.
The proof follows immediately from the facts that 2 is finite-dimensional (by the definition of a normal point) and is non-degenerate (by Theorems 2.20
and 2.24) taken in conjunction with 1, Corollary 7.18. We note that in the conditions of the corollary it is not excluded, for example, that X E p(A), and in that case 2),(A) = (0).
5 Now let .
= H. be a Pontryagin space. We investigate the structure of root lineals of 7r-dissipative operators.
Theorem 2.26: If A is a closed 7r-dissipative operator and a= « E a,, (A), then the root lineal 91, (A) is expressible in the form -Va (A) = /l "a [+] Afa, where dim .'U,. < co, AI V. C V,,,, ,A, C Ker (A - al) and a is a non-degenerate
subspace or, in particular, /tfa = (0). Moreover, if d1(a), d2(a), . . ., d,,, (a) are the orders of the elementary divisors of the operator A I . t"a, then
z
2
a=&Eap(A) i=1
di(a)J + Z
Im X>0
dim Y,\(A) < x.
(2.5)
First of all we note that, by virtue of Theorem 2.6, A can be assumed, without loss of generality, to be a maximal closed 7r-dissipative operator, and
therefore A = II,,. Since every non-negative subspace in H,r has a dimension not exceeding x
(see 1.9.2) it suffices in proving our theorem to verify that the indicated decomposition of -Ta (A) exists, and that in every ,/Y, there is a neutral subspace, invariant relative to A, of dimension Ei'=1 [21 d1(a)], and then to use Corollaries 2.18 and 2.22. We consider the decomposition of Ker(A - al) into its isotropic subspace
91, and a non-degenerate subspace Jfa: Ker(A - aI) = Y Q' [+] ., .. From Theorem 1.9.9. it follows that lfa is a projectionally complete subspace, and therefore H,, = ffa [+] H ,, where 1 1,', = lfa1) is again a Pontryagin space with
x 1 < x. Using Corollary 2.16 and Proposition 1.11 we obtain that M,
is
invariant relative to the operator A, and therefore it sufficies for us to prove the theorem under the assumption that Ker(A - aI) = 2,, i.e., Ker(A - al) is a neutral (and therefore a finite-dimensional) subspace. We introduce the
§2 Dissipative operators
101
notation 2° = Ker(A - al )° (p = 1, 2.... ). Since Y,, is finite-dimensional, it follows that all the Ya (p = 1, 2, ...) are finite-dimensional. We verify that, if df°1, d¢°),...,d?)(d;°) 5 d;°) when i 5 j) are the orders of the
elementary divisors of the operator A I Y°', then a neutral subspace of dimension E;=, [Z d;°)] exists in g'°. From Theorem 2.15 it follows immediately that, if (A - al )'x = 0, then
(A - aI)°xE 9) (A` - aI )1-° and
(A` - aI)'-°(A - aI)°x = 0 (p = [!],..., l - 1). z Let the vectors x, be such that d,")
1f' = Lin(x;,(A - al)x;,...,(A- al)d%"-'x;lfr then
2=Lin((A-al)[(dm'+;)/2)x;,...,(A-al)d;°'-Ixdi
is a neutral subspace. For, suppose that rI
2 11 °+ d`
q; S d; p1 - 1
and
rd; °2+ 1]qjdJ1_1
(i 5 j)
ThLLLlllen
qj
d; °- q;, and therefore
L
[(A - «1)9rx1, (A - aI)9"x;] = (A-al)d;°'-q;(A-aI)4i [(A-a1)qx, _ [(A`-a1)ds"-q,(A-aI)q;x
-(d;'"-q;)z;]
(A-al)q,-(d%°'-q;)xi]=0
Remark 2.27: The decomposition (A) _' I [+].ill can be chosen in various ways, but the number of non-prime elementary divisors of a given
order r of the operator A I I', is an invariant of the operator A and the number a for any choice of e/1"a, because it is the same as the number of elementary divisors of order r - 1 of the operator generated by the operator A
in the factor-space Y. (A)/Ker(A - aI), and the latter is finite-dimensional and does not depend on the choice of Corollary 2.28: For a closed ir-dissipative operator A in n. all the root subspaces corresponding to the real eigen values are, except for not more than x of them, negative eigen-subspaces. Let
Jul be the set of those real eigenvalues for which there is in
Ker(A - al) at least one non-negative vector. It follows from Theorem 2.26 that, in particular, all those a for which Y. (A) ;4 Ker(A - al) enter into this set, and from Corollary 2.18 it follows that the set {al is finite and consists of not more than x elements.
102
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
Corollary 2.29: Let . be a W(")-space (i.e., in accordance with Proposition 1.9.15, it is assumed that the set a/w (x) fl [0, oo) contains precisely x (< oo ) (taking algebriac multiplicity into account) eigenvalues, let `?W,) =4w"1, and
let A be a W(`)-dissipative operator defined on J(. Then the assertion of Theorem 2.26 holds for the operator A. Let dim Ker W(")= xo(<x). By Lemma 2.19 A Ker WYYY C Ker and therefore in the Pontryagin '/Ker Wt"t (see Corollary 1.9.16) the zr-dissipative operator A^ generated by the operator A is properly defined by the formula (1.11). Theorem 2.26 holds for the operator A". It is not difficult to see that the left-hand side of the inequality (2.5), which was set up for the operator A exceeds the corresponding sum for the operator A" by not more than xo units, and therefore the assertion of Theorem 2.26 is true also for the operator A.
6
Definition 2.30: An operator A is said to be strictly J-dissipative if
Im [Ax, x] > 0 for (6 6 )x E CAA,
and, in particular, to be uniformly J-
dissipative if there is a constant yA > 0 such that Im [Ax, x] 3 yA x (xE 2A) Theorem 2.31: Every closed strictly or uniformly J-dissipative operator A can be extended into a maximal closed J-dissipative operator A which is also respectively strictly or uniformly J-dissipative, and moreover the latter can be chosen so that yd = yA.
We consider the graph IPA of the operator A. In both cases, by the form (1.4) it is a positive subspace, and therefore (see Remark 1.8.9) I'A admits extension into a maximal non-negative subspace which is positive. Since the latter is non-degenerate it is the graph of a maximal J-dissipative operator A (see Theorem 2.6) and it is, moreover, strictly J-dissipative. Now let A be a closed uniformly J-dissipative operator. This is equivalent to saying that the closed operator A - iyA J is J dissipative. By Theorem 2.6 this
operator admits extension into a closed maximalJ-dissipative operator
A J, and therefore the operator A = A - iyA J + iyA J is a maximal J-dissipative
operator
and
it
is
uniformly
J-dissipative:
IM [Ax, x] '> yA II x 1; 2 (x E VA).
To conclude this paragraph we note that 2.32 Strictly J-dissipative operators have no real eigenvalues. All real points are points of regular type of closed uniformly J-dissipative operators; they are
regular points of these operators if and only if the latter are maximal J-dissipative operators. If A is a strictly J-dissipative operator and (A - Xol)xo = 0 when X0 = Xo Ci then Im [Axo, xo] = 0, which, by definition, is possible only when xo = 0, i.e., Xo0ap(A).
§2 Dissipative operators
103
Now let A be
a uniformly J-dissipative closed operator. Since (A-Xo1)x1I >Im[(A-Xol)x,x]/II x11=Im[Ax, x]/Hx1I >, -yAIIxII when
II
(0 ;x-' )x E 1A and Xo = Xo, so Xo is a point of regular type for the operator A. If
also A is a maximal J-dissipative operator, then from Corollary 2.17 we have Xo E p(A). Conversely, if A is a uniformly J-dissipative operator and Xo = 4o E p(A), then A is a maximal J-dissipative operator, since otherwise by Theorem 2.31 it would admit uniformly J-dissipative extensions for which the point Xo would already be an eigenvalue-we have obtained a contradiction.
Exercises and problems 1
Show that in Theorem 2.9 the condition Ye+ C 9A is essential: give an example of a maximal J-dissipative operator with p(A) = 0; give an example of a non-maximal J-dissipative operator with C' C p(A).
a) b) 2
3
(0, oo) Let Y be a G(')-space, i.e. (cf. Proposition 1.9.15) 0)F op(G(") and consists of x < oo eigenvalues (taking multiplicity into account), and let A be a Gt" -dissipative operator. Prove that then dim Lin[?1,(A) I Im X > 0) < x.
Give an example of a
maximal a-dissipative with
a(A) n c- ;e o.
a(A) fl c,
0 and
4
Let .0 be a G (')-space and let A be a G (x )-dissipative operator in ,Y. Prove that dim (A - aI )91" (A) < oo when a = a.
5
Under the conditions of Example 4 determine the orders 61(a), 62(a), ..., 6,(a) of
the elementary divisors of A in (A - al ).P" (A); we shall call the numbers dt") = S;(a) + 1 (j = 1, 2, ..., r) the orders of the non-prime (d; 2) elementary divisors of the operator corresponding to a. Prove that the formula (2.5) holds for A (cf. Usvyatsova [1]). 6
Let WT = [ [ Tx, x] I x E 1IT). Prove that either WT = C, or W = IR, or WT is a
certain angle with its vertex at the origin of coordinates and its aperture Or < a. Verify that if BT 5 Tr then there is a number ,p E fR such that e''0 T is a J-dissipative
operator (cf. [XI]). 7
Prove that any (maximal) J-dissipative operator can be appoximated in the uniform operator topology by (maximal) uniformly J-dissipative operators.
8
Give an example of a strictly J-dissipative operator which does not admit closure.
9
Prove that every (including those not densely defined and non-closed ones) uniformly J-dissipative operator admits maximal closed uniformly J-dissipative extensions.
10
Prove that if A is a closed maximal 7r-dissipate operator and YW' C VA, then A 12
is an Au-completely continuous operator. I1
Prove that if A is a W-dissipative operator and X E ap(A) and Axo = Xxo, Ax1 = Xx1 + xo, then xo is an isotropic vector in Ker(A - XI). Hint: Use the fact that the form Im [Ax, x] is non-negative on vectors X E CIA-
12
Prove that in Corollary 2.28 the requirement that the operator A be closed can be dropped. Hint: Use instead of Theorem 2.26 the result of Exercise 11.
104 13
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric Let . V' = _Y' U .
- be a J-space, let Y' E .//', and let K-, - be its angular
operator. Prove that if .SP' and 99' i'J have no infinite-dimensional uniformly definite subspaces, then K,- is a 4-operator, and if K,-- = T I K.-- I is its polar representation, then T is also a 4)-operator, and ind T = ind K., - (Azizov).
Hint: Use the result of Exercise 17 to §8 in Chapter 1, the definition of a 4)-operator in §2.2, and Theorem 2.11. 14
Suppose ,y'' E //' and 99' 1 Ll contain no infinite-dimensional uniformly definite subspaces. Let K, - = T I K, - I be the polar representation of the angular operator K,- of the subspace 2', let I K,- I = I' + S, S E .9. (see Exercise 17 to §8, Chapter
1), and let KM- be the angular operator of the uniformly positive subspace M' E It'. Then K,, - - KM- is a 4)-operator, and ind(K. - - KM-) = ind K,, (Azizov).
Hint: Verify that K,- - KM- = TV+ S1, where V:.,Y' .,Y' is a linear homeomorphism, and S, E y.; use the result of Exercise 13 and Theorem 2.11.
§3
Hermitian, symmetric, and self-adjoint operators
Definition 3.1: An operator A operating in a W-space .YP is said to be W-Hermitian if Im [Ax, x] = 0 for all x E 1A. In particular, an operator A is W-symmetric if also A = . f, and it is a maximal W-symmetric operator if it does not admit W-symmetric extensions A D A, A ;4 A. From Exercise 1 to § 1, Chapter I it follows that an operator A is W-Hermitian if and only if [Ax, y] = [x, Ay] for all x, y E A. Therefore, if A is an operator in a G-space (we recall that 0 a, (G)), then its G-symmetry is equivalent to the inclusion A C A`. Since the operator A` is closed (see 1.5), a G-symmetric operator admits closure. 1
Definition 3.2: A G-symmetric operator is said to be G-self adjoint if
A=A`. We now return to the study of the properties of W-Hermitian operators. It follows from Zorn's lemma that 3.3 Every W-symmetric operator admits extension into a maximal W-symmetric operator. From definitions 2.1 and 3.1 we obtain immediately
3.4 An operator A is W-Hermitian if and only if A and (-A) are simultaneously W-dissipative.
From now on to simplify the discussion we shall again, as in § 1 and §2, speak mainly about operators acting in a J-space .Y, although many of the results are also valid in the case of more general spaces. It follows from Proposition 3.4 that
3.5 An operator A is J-Hermitian if and only if its graph is neutral in the metric (1.4) Moreover, the following holds:
§3 Hermitian, symmetric, and self-adjoint operators
105
3.6 A J-symmetric operator A is maximal if and only if rA is a maximal semi-definite subspace in ,Yr = ,Y x W. In addition, J-self-adjointness of the operator A is equivalent to the neutral subspace rA being hyper-maximal. This goemetrical proposition is a consequence of Propositions 3.5, 2.4, and Proposition 4.1 in Chapter 1; it can be rephrased in terms of operators thus:
3.7 A J-symmetric operator A is maximal if and only if at least one of the
operators A or (-A) is a maximal J-dissipative operator. Moreover, Jselfadjointness of the operator A is equivalent to A and (-A) being simultaneously maximal J-dissipative operators. Propositions 3.4-3.7 enable a number of the assertions in §2 to be made more precise for J-symmetric operators. Thus, for example, we have Corollary 3.8 (cf 2.3): J-symmetry (maximal J-symmetry, J-selfadjointness) of an operator A is equivalent to each of the operators JA and AJ being symmetric (maximal symmetric, self-adjoint).
Corollary 3.9 (cf. Theorem 2.9): Let A be a J-symmetric operator and .0+ C VA. Then the following assertions are equivalent: a) b)
A is a maximal J-symmetric operator; A22 is a maximal symmetric operator in
c)
[XIImXI
d)
((X I Im X > II AP+ III C p (A )) V ((X I - Im X>211 AP+ II) C p (A ))
Corollary 3.10 (cf. Corollary 2.12):
-;
0;
If A is a maximal J-symmetric J-
selfadjoint operator, if ,+ C 9A, and if A12 is an A22-completely continuous
operator, then C+ C p(A) or C- C OS(A) (C+ U C- C p(A)). Moreover, if C+ C p(A) (C- C p(A)), then C-(C+) consists of points of regular type and not more than a countable set of eigenvalues of finite algebraic multiplicity of the operator A. In view of Corollary 2.12 explanation is needed only for the last assertion.
But this follows from the J-symmetry of the operator A and Theorem 1.16.
Corollary 3.11 (cf. Theorem 2.21): Suppose .0 is a W-space, A a W-symmetric operator, and a is its bounded spectral set: a C C+ or a C C-. Then PQ.Y( is a neutral subspace.
Corollary 3.12 (cf. Theorem 2.24): Suppose A is a J-self-adjoint operator, and a + is a bounded spectral set, a + C C+ . Then a - = a '*(C C-) is also a spectral set of the operator A, the projector P. - uo - is J-self adjoint (we shall also call such a projector J-orthogonal ), and therefore PQ - uQ - -W is a projectionally complete subspace. Moreover, if at (C C+) is another spectral a l ' * , then P. - uo set o f the operator A and a + fl al, = 01 and al[1) PQ, uo,-Jf.
106
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
By Theorem 1.16 the spectrum of a J-self-adjoint operator is symmetric about the real axis, and therefore a - = a+ * is a spectral set of the operator A together with a+ . Let ro - u o - = FQ - U I, -, where IQ- is a contour consisting of
regular point of the operator A and surrounding a +, FQ- C C, and rQ- = IF,*-. Then PCO -U.
I -2iri1
re--
_- 1 2iri r
A
-XIdx)
=2ri
r
A - XI' dX=P
As for the projectional completeness of the subspace PQ - uQ -.W, it follows
from the fact that the whole space splits up into the sum of the subspaces and (I - Pa - uQ -), ' which are J-orthogonal to one another: Po - uQ
(PQ'uo-X, (I- PQ'u.-)Y] _ (x, Pv'uv-(1- PQ-uQ-)Y] =0. The J-orthogonality of P. - uQ -.1( and PQ - u o -.W follows from Theorem 2.20.
Remark 3.13: In Corollary 3.12 we could, formally, weaken the conditions on the operator A by premising, not that it is J-self-adjoint, but only that it is J-symmetric and o+, a- satisfy the conditions of Theorem 2.24. But this relaxation is only formal, since from the conditions on the operator A -it follows that it has at least one pair of non-real regular points Xo and Xo symmetrically situated relative to the real axis. If A ;e Ac, then Xo, Xo E op (A`); but this contradicts assertion b) of Theorem 1.16, and therefore A is a J-selfadjoint operator. Corollary 3.14 (cf. Corollary 2.22): Suppose that A is the set of eigenvalues of a J-symmetric operator A and A fl A* = 0. Then C Lin (2x(A)}x E A is a neutral subspace.
Corollary 2.18 has to be used in proving this assertion. Corollary 3.15 (cf. Corollary 2.23): Let A be a 7r-selfadjoint operator in II,. Then its non-real spectrum consists of not more than 2x (taking multiplicity into account) normal eigenvalues situated symmetrically about the real axis.
The symmetry of the spectrum of a J-self-adjoint, and, in particular, of a ar-self-adjoint, operator follows from Theorem 1.16 (cf. Exercise 1 below). It remains only to use Proposition 3.7 and Corollary 2.23. Corollary 3.16:
Let A be a ir-self-adjoint operator. Then o,(A) = 0.
This assertion follows from Corollaries 2.17 and 3.15.
§3 Hermitian, symmetric, and self-adjoint operators
107
Definition 3.17: We shall say that a J-symmetric operator A is semibounded below if yA = inf [ [Ax, x]l(x, x) I (00 )x E cA) > - oo. In particular, 2
an operator A is said to be J-non-negative (A "0), J -positive (A>0), or uniformly J-positive (A)' 0) if, respectively, [Ax, x] > 0 when x E c/A, [Ax, x] > 0 when 0 ;d xE CAA (in both cases 'A >, 0), or [Ax, x] 3 y(x, x) when x E c/A and for some y > 0 (i.e, yA > 0).
In particular, if J = I then Definition 3.17 repeats the corresponding definitions for symmetric operators and in this case we shall omit the symbol
'I' above the signs > , > ,
.
Definition 3.18: Let A and B be J-symmetric operators and let 1A C VB.
We shall say that A>B if A - B'O. Theorem 3.19: Every J-non-negative operator can be extended into a J-self-adjoint J-non-negative operator. Before proving this theorem we shall deal with some auxiliary propositions and we introduce the concept of a 'quasi-inverse' operator. Let T be a densely defined operator. Then we shall say that the operator
T(-')= Q(T l.N?T*)-'P
(3.1)
where P is the orthoprojector on to 4T, and Q is the orthoprojector on to .T*, is quasi-inverse to T. Lemma 3.20: If A and B are bounded operators with CIA = c/B =,W and A *A <, B* B, then A = KB, where K = AB(-') and therefore 11 K II < 1.
KB = AB(-')B = AB(-')B = A. Let be a Hilbert space, ,? a subspace of it, .Y = 2 O+ 21, and relative to this decomposition let a bounded operator T be expressed in matrix form: T = II T,, 11 ?i=1. For the operator T to be non-negative (respectively, positive, uniformly positive) it is necessary and sufficient that the operators Tl1 and T22 be non-negative (respectively, positive, uniformly positive), that T21 = Tie, and that there be an operator S12: 2'1 2', when 11 S12X 11 < 11 x 11 (respectively, V's,: = 21, S12 < 1 (B34 )xE WT2=2; S12 < 1), Ker T22 C Ker S12, JPs,_ C 4?T,,, such that Lemma 3.21:
T12 = Till S12 '222
Sufficiency: Let T11, T22, and S12 be the operators indicated in the theorem. Then Tii`S12Tzi` - Ti, T12 7 ' = II T11 Tz
T22
I
I
-
II
Tu2Si Tii2
T22
11
-
(3.2) I
I T012
Toe 2 I I
S2
12 I
I
I I TO 2
TOz e
ll
108
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
It can be verified immediately that the middle factor is a non-negative operator if II S12 II < 1, and is uniformly positive if II S12 II < 1. This remark enables the
non-negativeness and uniform positiveness of T in the cases indicated in the theorem to be proved easily. Now let T11>0, T22>0, and II S12Tzzzx II < II Tzz2xIl for all (0
)xE tel. Since lrz2z = L, so
II S12
1, i.e., T > 0. But if we had
x E Ker T, x = x1 + x2 where x1 E 2, x2 E Y l , then this, taking (3.2) into account, would imply the system of equalities Tii 2x1 + S12 7iz2x2 = 8,
{S12Tiizxl+Tzzzxz
( 3.3 )
=0.
Hence we obtain II S12T2' 2xz II = II Ti'( x1 II > IIS zT1i2x1 II = II
T21/2X2 II
which is possible only when x2 = 0. But then we conclude from the first inequality in (3.3) that Tii 2x1 = 0, and since T11 is positive we have xl = 0, i.e.,
x = 0. Therefore T > 0. Necessity: The equality T21 = T 2 follows from the self-adjointness of T. Since (T;;, xi) = (Tx,, x;), i = 1, 2, xi E 2'1X2 E 911, the conclusion regarding T11 and T12 is correct.
Now let T T. 0. Then it is clear that the operator S12 = Ti11nT12Tz21'2 satisfies our requirements. But if T is an arbitrary non-negative operator, then T + (1/n)I -> 0, n = 1, 2, ..., and therefore it follows from what has been proved above that there are operators Siz ) with 1 vz z / (n) II Siz II<1, n T12T11+IS}T22+J)1/
n // Without less of generality (see 1.8.20) we may suppose that, as n - 00, the sequence Siz converges in the weak operator topology to an operator Sit with II S1'2II < 1, and therefore T12 = Tii2S1'2Tz'2. We take as the operator S12 S12 =QSi2P= (Tii2)(-')T12(Tzz2)( '), where P is the ortho-projector on to 4T12 and Q is the orthoprojector on to, ?r,,. The further verification of the properties of the operator S12 presents no difficulties. Suppose that A is a Hermitian operator defined on a subspace VA of a Hilbert space .W = VA O QOA and that relative to this decomposition A can be expressed in the form of the `vector'
A=
A12
A22
This operator will be a contraction if and only if A22 is a contraction and A iA 12 < 12 - A 22 (here 12 = I I r/'A), or, what is equivalent by Lemma 3.20, A12=K(I2-A22)"2,
22)1/2] where K=A12[(12-Az2)'
and
11 K11 < 1.
A Hermitian operator A always admits selfadjoint extensions A on to the
§3 Hermitian, symmetric, and self-adjoint operators
109
whole space ', and these extensions can be expressed in matrix form A=
X* A12, A12
(3.4)
X11 = Xi,.
A22
If A is a contraction then it is easily seen that, if I, = I I
Ail = -I,+A12(12+A22)
=
LAAn,
-Ii+K(12-A22)K*
I1 - K(12 + A22)K* = I, - A12(lu - A22)t ' A22 = Ail
(3.5)
Also, A is a selfadjoint contraction if and only if -I < A < I. By Lemma 3.21 this is equivalent to the system
-I,<X1,11, A,2 = (I, - X,,)v2S,,2(12 - A22) 1/2' A 12 = (11 + X11)
1,12S1
(12 + A22)
1/2,
where Sit and S,"2 have the properties indicated in Lemma 3.21. Since A,2 = K(I2 - A 22) "2, this system can be rewritten in the equivalent form
-11<X11<1,, K(I2+A22)1/2=(12- X,,)1i2S1,2,
K(I2 -
A22)
1/2
= (1, + X,, )1r2S1n2,
which in turn is equivalent to the inequalities
- I, + K(I2 - A22)K* < X,, < I, - K(I2 + A22)K*. Taking (3.5) into account we have proved Lemma 3.22:
If the operator _ A 12 A-IA22
is a Hermitian contraction then it admits extension into a self-adjoint operator
A of the form (3.4), and the latter will be a contraction if and only if Ai1<X11
By Corollary 2.5 we can suppose without loss
of generality that A is a closed J-non-negative operator. Since A being a J-non-negative J-symmetric operator is equivalent to iA being a J-dissipative operator, the graph r,A of the latter is a non-negative subspace in the space 4Yr (see §1.2), and its angular operator K;A = (V, - JVJ), where V= (J- JA)(I+ JA)-1 (see (1.8), (1.9)) is a contraction since JA > 0, so V is a Hermitian contraction. In accordance with Lemma 3.22 V has a self-adjoint extension V with 1117 II < 1 which generates the angular operator (v, - JVJ) of a certain maximal non-negative subspace in ,Yr that is an extension of rjA and which is, by Proposition 2.4, the graph of a certain
110
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
J-dissipative operator
iA.
It
follows
once again
from (1.9)
that
V = (I - JA) (I + JA)-', or what is equivalent,
A= J(I-V)(I+Vy'= J-2J(I+ V) From Propositions 1.6, 1.9 and Corollary 3.8 it follows at once that A is a
J-selfadjoint operator, and this combined with the fact that iA is a Jdissipative operator proves the theorem.
Corollary 3.23: Let A be a J-symmetric operator semi-bounded below: [Ax, x] > yA(x, x) (x E !YA). Then it has J-selfadjoint extensions A semi-bounded below such that yA = yA. For any such extension
OEa(A-yAJ). The existence of the required A is equivalent to asserting the existence of J-selfadjoint J-non-negative extensions A y J of the J-non-negative operator A - J. Such as extension exists by virtue of Theorem 3.19 and
therefore A = A -- yJA J is the required ^exten ion of the operator A. For, [Ax, x] _ [A - -yA Jx + -yA Jx, x] _ [A - yA Jx, x] + yA (x, x) > yA (x, x),
and therefore -yi > yA. On the other hand, since A J A, so yA
'yA, i.e., yA= 'AWe now verify that 0 E a(A - yA J). Suppose the contrary: 0 E p(A - yA J). But then the J-non-negative operator A - -yA J would be uniformly J-positive, and therefore we would have yA> yA, which is impossible.
Corollary 3.24:
Let A be a J-symmetric operator. Then A > 0 (A > 0, A ' 0)
if and only if iA is a J-dissipative (strictly J-dissipative, uniformly Jdissipative) operator. Moreover, iA is a maximal J-dissipative operator if and
only if A = A`>0. The first part of the assertion is trivial, and we have already used it in proving Theorem 3.19. If A = A`>0, then (see the proof of Theorem 3.19) F,A
is a maximal non-negative subspace in .0 and therefore by 2.4 iA is a maximal J-dissipative operator. Conversely, if iA is a maximal J-dissipative operator, then A has no non-trivial J-non-negative extensions, and therefore it follows by Theorem 3.19 that A = A`>0
If Then a p (A) fl (C+ U C -) = 0. Corollary 3.25: Let A > 0. A = J X E ap (A) l X > 0(<0)1, then 1), (A) = Ker(A - XI) and Lin(Ker (A - XI) I X E A} C .,P+ + U (0) (P- -U 10}). If X= 0, then, generally speaking, 1o(A) * Ker A, but the lengths of the Jordan chains of the operator A I 2o(A) do not exceed 2.
By virtue of Theorem 3.19 we can suppose that A = A `> 0. If X E ap(A) fl (C+ U C
and Axo = Xoxo (xo ;d 0), then from Corollary 3.14 we
obtain that [xo, xo] = 0, and therefore [Axo, xo] = 0. Using the CauchyBunyakovski inequality, we obtain Axo = 0-a contradiction. Essentially we
§3 Hermitian, symmetric, and self-adjoint operators
III
have proved that the condition xo E Ker(A - XoI) fl j° implies X0 = 0. Therefore, if (0? ) X E vp(A) and Ax= Xx, then [x, x] = [Ax, x]/X, hence Ker(A - XI) C .40 + + U (0)(,?- - U 10)) if X > 0 (<0), and therefore (see
§2, Problem 11) Y,\(A) = Ker(A - XI) when X 34 0. The assertion about Lin(Ker(A - XI))xEn follows immediately from Corollary 2.18. Now let Axe = x,, Ax, = xo, Axo = 0, i.e., A3x2 = 0. Then 0 = [A 3x2, x2l = [AAx2, Ax2l, and therefore x, = Axe E Ker A, i.e., xo = 0.
If A = AA `> 0, then Q,(A) = 0.
Corollary 3.26:
By Corollary 2.17 in combination with Proposition 3.7 A(= A') has no real points in the residual spectrum. If X 3;6 X belonged to o,(A), then by Theorem 1.16 X E ap(A), contrary to Corollary 3.25. We now investigate the spectrum of a J-self-adjoint J-non-negative operator. Theorem 3.27:
Let A = A0 and p(A) ;d 0. Then C+ U C- C p(A).
We assume at first that A is a continuous operator. Let lo # Xo E a(A). From Corollaries 3.25 and 3.26 we conclude that Xo E ac (A), and therefore there is a sequence Ix)' of normalized vectors (II x,, 1, n = 1, 2, . . .) such that (A - XoI )xn
when n
0
ao.
(3.6)
Since
RA - XoI )x,,, xn] _ RA - Re XoI )xn, x,r] - i Im Xo [xn, xn] and Im Xo ;4 0, so [x,,, xn]
0 as n --' oo, which implies that
[Axn, xn]- 0 From the continuity of the operator A and with it also that of
(JA)1,'2
it
follows that
Axn = J(JA)112(JA)1 2xn
0
(n --* oo).
Taking account of (3.6) we obtain xn 0 (n oo)-a contradiction. Now let A be an arbitrary J-self-adjoint J-negative operator, and (µo pd)µo E p (A ). It follows from Corollary 3.12 that µo E p SA ), and so the
operator B = (A - µol)-'A(A - µoI)-' is bounded, B = B`>0. By what has been proved above, a(B) fl (C+ U C-) = 0. A well-known theorem of Dunford (see [VII]) about the mapping of the spectrum asserts that a(B) = {µIµ = X/(X - µ0)(X - µ.o), X E v(A)). It follows from X0 Xo and Xo I # I µo I that the number Xo/(X - µo)(X - µo) is not real, and so Xo E p(A), i.e.,
(C+UC-)\[XIIXI=IµoI)Cp(A). We now take (µ,
)µ, E p (A) with 114 1
1
;6 1 µo I ,
(3.7)
and carrying out a similar
112
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
argument we obtain (C+ U C )\ I X I I X I = l
141 1
1 C p(A ).
(3.8)
It follows from (3.7) and (3.8) that C+ U C- C p(A). Corollary 3.28: Let A be a ir-self-adjoint ir-non-negative operator in H,. Then C+ U C- C p(A), and the set a(A) fl (0, co) consists of not more than x normal eigenvalues (taking multiplicating into account). From Corollary 3.15 and Theorem 3.27 we conclude that C+ U C- C p(A). The second part follows from Corollaries 3.24 and 2.23.
Corollary 3.29:
Let A = A` 0. Then C+ U C- U (0) C p(A). If a+ (respec-
tively a-) is a bounded spectral set of the operator A, a+ C (0, 0°) (resp. a- C (- oo, 0)), and P. - (resp. P. -) is the corresponding Rim projector (2.4), then PQ-.W (resp. PQ-.) is a uniformly positive (resp. uniformly negative)
subspace. In particular if A =A c 0 is a bounded operator, then with a+ = a(A) fl (0, oo), a- = a(A) fl (- oo, 0) the subspaces PQ- and PQ-Jf are J-orthogonal to one another, are maximal uniformly definite, and = PQ-.Yf [+] PQ -X.
.
From Corollary 3.24 and Proposition 2.32 we conclude that 0 E p(A). We use Theorem 3.27 and obtain the inclusion C+ U C- U (0) C p(A). From Corollary 3.12 follows the projectional completeness of the subspaces PQand PQ -., and from Theorem 2.21 and Corollary 3.24, their uniform definiteness with the proper signs. If A is a bounded operator, then, since a(A) = a+ U a-, a+ fl a- = 0, we have by Theorem 3.20 PQ- + P,,- = I, i.e., using Corollary 3.12, ye = PQ-,W [+] PQ- . The maximal uniform definiteness of PO .Y( and PQ now follows from Proposition 1.25 in Chapter 1. 3
Here we present some examples which illustrate the preciseness of the
conditions imposed on the operators in the theorems in §§2-3, and the essential difference of the indefinite case from the finite one.
Example 3.30: A Hermitian operator which does not admit closed dissipative extensions (cf. Theorems 2.6, 2.31, 3.19). Let _+ O+ hl- be an infinite-dimensional space with -W+ = Lin (e+ }, 11 e+ 11 = 1, and let p be a discontinuous linear functional defined on-. Then the operator A: x- - ,p(x- )e+ with cA = ye- is Hermitian, since
(Ax-, x-) = 0. If it admitted a dissipative extension A, then #A= .W, and therefore (see Corollary 2.5) A' would be a closed operator. But then by Banach's Theorem A would be a continuous operator, which contradicts the discontinuity of gyp.
§3 Hermitian, symmetric, and self-adjoint operators
113
The operator constructed also serves as an example of a Hermitian operator which does not admit closure and therefore the closure rA in ,Yr of its graph rA would no longer be the graph of an operator. Example 3.31: An operator A = A` .O with a(A) = C (cf. Lemma 2.8, and Theorems 2.9 and 3.27).
Let ,W be a J-space and (2t, 22) a maximal dual pair (see 1. §10.1) of definite (but not uniformly definite) subspaces 9t and 22. We define an operator A: -QA = -T1 [-+] 92, A(xt + x2) = xi - X2, xt E 2l, x2 E 92. By Lemma
,e, and since A9A C 9A, so a(A) =C. It remains to verify that A =A`>0. Since [A(xl + x2), xt + x2] = [xi, xl] - [x2, x2] > 0 when xl + x2 * B, so A .0. Now let y E 9A° and ([A (xi + x2), y] = ) [xt - x2, y] = [xi + x2, z] for all 7.7 and Corollary 7.17 in Chapter 1 we have 1A = Ml, 9A
xl + x2 E I1A and, in particular, [xl, y] = [xl,z] and [ - x2, y] _ [x2, z]. Therefore y + z E 21 and y - z E 22, and therefore y E 9A( = 91 [+] 92), i.e.,
VA' C 9A, which proves that A is J-selfadjoint. We note also that A 2 = I l !JA, and Pi = 2'(I+ A) and P2 = z (I - A) are the J-orthogonal projectors from !2A on to 2t and 22 respectively, and also Pt >, 0, P2 >, 0, and
a(Pi) = a(P2) = C. Example 3.32: A 7r-self-adjoint operator, in a finite-dimensional Pontryagin space II,,, for which the inequality (2.5) becomes an equality (cf 2.26).
Let .*'= Lin(eo, el, e2) be a unitary three-dimensional space with an orthonormalized basis (eo, el, e2). Each vector x E ,' is expressed in the form x = oeo + Elel + E2e2 Let y = floeo + n1e1 + q2e2 We consider the indefinite form [x, y] = - 6n2 - tl' l - E2fio It is generated by the operator J:
Jeo = - e2, Jet = - el, Je2 = - eo, the eigenvalues of which are X = -1 and X = 1 with multiplicities 2 and 1 respectively, i.e., .' =1I with x = 1. We a linear operator A; Aeo = 0, Atte1 = eo, Ae2 = el. Since [Ax, x] = [ leo + E2e1, Soeo + tie, + S2e2] = tltt2 + E1E2 is a real number we
define
have A = A`. If follows from the definition of the operator A that it has a single eigenvalue X = 0 and a single Jordan chain of length d = 2x + 1 (= 3). Therefore [d/2] = x (=1). Example 3.33: A closed positive operator A which does not admit positive selfadjoint extensions..
We presuppose that A is a closed, non-selfadjoint, positive operator with finite-dimensional deficiency numbers (i.e., dim ;?IA_ar < oo and dim A _ v < oo when X ;;d X), then X = 0 is a point of regular type for A, and that yA = 0. By Corollary 3.23 for any selfadjoint non-negative extension A of this operator 0Ea(A). From [I] (see also Azizov, I. Iokhvidov, and V. Shtrauss [ 1 ] ) it follows that .mod = RA and therefore, whenever 0 ¢ p (A-), we have 0 E ap(A), i.e., no non-negative selfadjoint extension of the operator A with the properties indicated above will be positive.
114
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
We give an actual example of such an operator A. = Lin(e) O L2(0, 1) with II e II = 1. We consider in L2(0, 1) the Let selfadjoint contraction V22: V22x(t) = tx(t) (x(t) E L2(0, 1)) and the contraction K: L2(0, 1) Lin{e}, Kx(t) = to x(t) dt e. Then (see the proof of Lemma 3.22) the operator
v-
I K(122
will be a Hermitian contraction.
Since it follows from (I - V)x(t) = 0 that x(t) = 0, the operator
A: A(I- V)x(t)=(I+ V)x(t) will be properly defined on VA = (I - V )L2 (0, 1). Since
(A(I- V)x(t),(I- V)x(t)=((I+ V)x(t),(I- V)x(t)) (1 - t2)Ix(t)I2 dt-
=
('
(1
- t2)1,12x(t) dt
2
J0
is a positive number when x(t) ;d 0, it follows that A is a Hermitian operator. If we suppose that the vector z(t) _ Xe+ zl(t) (zi(t) E L2(0, 1)) is orthogonal to VA, then we obtain that 1
(1 - t)x(t)zl(t) dt - X
1
(1 - t2)1,'2x(t) dt = 0,
0
0
and therefore
zl(t) =
(I
t) 1,12
+ t)112 E L2(0, 1),
which is possible if X = 0, i.e., z(t) = 0. Thus,
A=
and A is a positive
operator. That A is closed follows from the facts that V is closed and
A=2(1-V)-'-I.
From Corollary 3.24 and taking Lemma 2.8 into account, we obtain that X = - 1 is a point of regular type for the operator V, i.e., IA = (I+ V )L2 (0, 1) is a subspace and therefore X = 0 is a point of regular type for the operator A. The finiteness of the deficiency numbers of the operator follows from the fact that RA = Lin(e+ [(1 - t)/(1 + t)] 1/2}, and therefore dim i1A = 1. It remains to verify that '(A = 0. To do this we observe that, if xp(t) E L2(0, 1) are chosen so that the sequence ((1 - t2) 112xp(t)} converges in L2(0, 1) to xo(t) = 1, then
limp- .(A(1- V)x,, (t), (I-V)xp(t))=0, but limn-..((I- V)xp(t), (I- V)xp(1))=2, and therefore inf((A(I - V)x(t), (I - V)x(t)) x(t)ELAO, 1)}=0. -yA = ((I - V)x(t), (I - V)x(t)) Example 3.34: A J-selfadjoint differential operator. Let W = L2 (- 1, 1). We introduce a J-metric into it by means of the
operator J: Jx(t) = x(- t) (cf. Remark 1.2.3 and Exercise 1 to 1.§3). We
§3 Hermitian, symmetric, and self-adjoins operators
115
consider the differential operator A = (d2/dt2) - 2(d/dt) + 1 with the boundary conditions x(- 1) = x(1), x'(- 1) = x'(1) and the maximal domain of definition in L2(- 1, 1). Since we have for twice-differentiable functions x(t) and y(t) with the given boundary conditions [Ax, y] =
(x"(t) - 2x' (t) + x(t))y(- t) dt
Ax(t)y(- t) dt =
1
J
1
= x' (t)y(- t) - 2x(t)y(- t)
1
+ f
1
1
x(t)y(- t) dt
f' x'(t)y'(-t)dt-2 f ' x(t)y'(-t)dt
+
1
=x(t)Y'(-t)
1 1
=
+ J1 x(t)(y" 1 - t) - 2y'(- t) + y(- t)) dt 1
x(- t)Ay(t) dt - [x, Ay],
f1
we see that A is a J-symmetric operator. The J-selfadjointness of this operator follows from the independence of the boundary conditions. It can be verified immediately that the eigenvalues of this operator are the numbers
(non-real if n P, 0) X" = - r2n2 + I - 2rni for n = 0, ±1, ±2,..., and 0 Y,,,,(A) = Ker(A - X"I) = Line").
Example 3.35: A completely continuous J-selfadjoint integral operator. Let a be a real function of bounded variation on [a, b], let w = Var a, and let K(s, t) be a Hermitian non-negative kernel (see IV, §3.3) which is continuous with respect to each of the variables and is bounded with respect to
the set of variables. Then the operator A: (Af)(s) = I b K(s, t)f(t) da(t) is completely continuous (see, e.g., I. Iokhvidov [2], I. Iokhvidov and Ektov [1]), and that it is J-non-negative in the_Krein space L. (a, b) follows from the fact that [Af, f] = IQ IQ K(s, t)f(s)f(t) do(s) da(t) > 0. In conclusion we indicate one way of constructing J-selfadjoint operators having a set of properties prescribed in advance.
j are the same as in Example 3.36: Let . if- be a J-space, where j 7fExample 1.3.9, let G = 0, and V be a unitary operator. Now let A 1, be a closed operator densely defined in Jr. ,Then the operator A: A(x1 + x2) = A1x1 + V*A1*Vx2,
where x, E 9A, X2 E V*JA; ,
is a J-selfadjoint operator. For,
A`=JA*J=
0
V
Al
V*
0
0
0
V
A
V*
0
0
0
V*A,V
V V* 0
0
0
0
V
V*A1* V
V*
0
Al
0
V*A1 V
= A.
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
116
In this example, by construction, A(-QA n ,Yi C ..Yi, i = 1, 2. We note also that the operator B given by the matrix
I
0
B1
B2
0
where B1 = Bi `, B2 = Bz*
will also be J-selfadjoint. Moreover, the conditions Bi > 0 (resp. B; > 0, B; . 0) are equivalent to the conditions B 0 (resp. B j 0, B)a 0). We leave the verification of these statements to the reader.
Exercises and problems 1
Reformulate Theorem 1.16 for a J-selfadjoint operator and by so doing obtain the `theorem on the symmetry of the spectrum of a J-selfadjoint operator' (Langer [21).
2 Prove that if A is a W-Hermitian operator and X, µ E ap(A) with X ;e µ, then
2a(A) [l] 'µ(A).
Hint: Carry out an argument similar to that used in solving Exercise 8 to §1. 3 Give examples of closed, densely defined operators A and B in the Krein space yP of Example 3.36 such that: a) the operator AA` does not admit closure: b)
JBB- [B).
4 Prove that if A is a closed operator in a Pontryagin space and VA = H., then AA` is a ir-selfadjoint operator (Kholevo [1]). 5
Let A be a closed operator in a J-space Ye with ;A = .Y['; suppose at least one of the following conditions holds: a) k+ C 9A; b) C c'A; c) A is a 4)-operator. Then (AA')' = AA` (Azizov). Hint: In cases a) and b) A = AP+A` + AP- A`, where one of the terms is bounded and the other is J-selfadjoint: in case c) prove that JAA, = .4Y and ind AA` = 0.
6 Construct an example of a bounded G-self-adjoint operator with a spectrum not symmetrical about the real axis (cf. Exercise 1) (Azizov [5]).
7 Prove that if A is a G-self-adjoint operator, and A = BG with B = B* and GB a closed operator, then a `theorem on the symmetry of the spectrum', similar to the theorem in Exercise 1, holds for the operator A in C\[0) (cf. with Exercise 6) (Azizov [2] ).
8 Give an example of a ,r-non-negative completely continuous operator which has not only eigenvalues but also principal vectors. Hint: Use the method indicated in Example 3.36 (cf. Example 3.32).
9 Use the method
in Example 3.36 to construct J-self-adjoint operators A1, A2, A, A4 such that: a) a,(A1) # 0, up(A1) fl C+ ;;d 0 (we recall that a,(A) = 0 always when A = A
b)
a(A2) = C (cf. Example 3.3); we recall that C+ fl c- C p(A) always when
A = A *); a,(A3) fl C+ ;e 0; d) the operator A4 has an infinite chain of eigenvectors and principal vectors.
c)
§4 Plus-operators
117
10 Let .YY be a J-space, and P a J-orthogonal projector on to a uniformly positive sub-space. Then there is an c = c(P) > 0 such that all J-orthogonal projector P' such that 11 P' - P 11 < e will map .Y on to the uniformly positive subspace P',J ' [XVIII]. Hint: First verify that for sufficiently small c the operator P' maps Rae homeomorphically on to P'.ie, and then prove that [P'Px, P'Px] > a Px 112 11211P'Px112forsome a>0). 11
iT = (P1 be a commutative family of J-orthoprojectors, with Rp C g+ for all P E J, and the subspace 2+ C ,? and P1+ C 1+ (P E .i). Prove that Let
C Lin (-Rp, 11+)pE., C ip+ (Langer [9]). Hint: Verify that the operator
(-Oil +i.+...+i"P'iP'2.,.Pii
P1,2, ,n = is
J-orthoprojector on to Lin(:i?.,.,lk=i e+ and 9'+ C 27+.
a
(1k=0,1)
(PkE., k=1,2,...,n),
that
.,, C
Then use the relation
[P1,2.....nx+y, P1.2,...,nx+Yl = [P1,2.....n(x+Y), Pi.2,...,n(x+Y)] + [(I - P1,2....,n)Y, V- P1,2,...,n)Y1,
which holds for all x, y E W.
§4
Plus-operators. J-non-contractive and J-bi-non-contractive operators
In this section the object of the investigation will be 'plus-operators', that is, operators whose existence depends on the indefiniteness of the spaces in which they operate, and also some sub-classes of plus-operators. 1
Let .', and W2 be a Wi-space and a W2-space with the ]I and [ , - ]2 respectively. A linear operator V. Yi W2 is
Definition 4.1: metrics
[
-
,
-
called a plus-operator if its dormain of definition contains positive vectors, and non-negative vectors are carried by it into non-negative vectors (gvn ++(.ie,) ;e 0; V(.?+(ae1) n !2 v) C +( 2)). Definition 4.2: An operator V. .i1 - W2 is said to be (W1, W2)-noncontractive if [ Vx, Vx]2 > [x, x] 1 for all x E 9 v. If follows from Definitions 4.1 and 4.2 that we may take as an example of a
plus-operator either an operator V having a non-negative range of values R,, or a V which is `collinear' with a (W1, W2)-non-contractive operator U, i.e., V = X U with X ;d 0. It turns out that, in fact, these examples exhaust the whole class of plus-operators.
Theorem 4.3: Let V: i1 - IY2 be a plus-operator. Then there is a number u > 0 such that [ Vx, Vx] 2 > µ [X, x] 1 (x E 9 v).
118
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
If J v is indefinite, then we are in the conditions of lemma I 1.35 and corollary I 1.36 ((t v = , V =) which it follows that [ V x , Vx] 2 > U [ x, x] ,
when A (V) < µ < µ+ (V ).
Since it follows from the Definition 4.1 of a plus-operator that {[Vx , Vx]2) [x.x]>o
[x,x]i
it is sufficient to put µ = µ+ (V) > 0. But if 9v is semi-definite, i.e., in our case non-negative (since by definition
of a plus-operator c v n io + + (.1) # 0), then we may again take as µ
µ+(V)=inf([Vx,Vx]2IxE!ty, [x,x],=11>0. Definition 4.4: A plus-operator V is said to be strict if µ+ ( V) > 0, and non-strict if µ+ (V) = 0. We note at once that the range of values of a non-strict plus-operator is a non-negative lineal. However, this property is not characteristic for the given class of operators. Indeed, let V = P+ be the orthoprojector on to the subspace W+ of the canonical decomposition W = W+ O+ W- of a Krein space W. It is clear that ?p _ Y C 40 +, but none the less µ+(P+) = 1 > 0.
Corollary 4.5:
A plus-operator V is strict if and only if it is collinear with
some (W1, W2)-non-contractive operator T: V= XT, and in this case
0
If V is a strict plus-operator, then by Theorem 4.3 the operator is a (W,, W2)-non-contractive operator. Conversely, if V= XT with X # 0 and T is a (W,, W2)-non-contractive
T= (1/X) V (0 < I X 12 <,a, (V))
operator, then [ Vx, Vx]2 > is a strict plus operator.
X2
[x, x],. Therefore
(V) > I X 12 (>O), i.e., V
It should be noted that plus-operators may be unbounded, and, what is more, they may not admit closure. As an example of such an operator it suffices to consider an operator V = ,p functional on .01, (B # )xo E JP+(. 2)
)xo, where
In this connection, criteria for the discontinuity of plus-operator acting in Krein spaces are of interest. We recall that the symbols P, (i = 1, 2) denote the canonical projectors (see 1.§3.1). Theorem 4.6: In order that a plus-operator V acting from a J,-space .°, into a J2-space W2 should be bounded (should admit closure) it is necessary and sufficient that the operator Pz V should be bounded (should admit closure).
It follows from Theorem 4.3 that the inequality 11 p2+ I'xllz+µlixlli>IIPZ VxII2 holds for arbitrary x E CA,,.
(4.1)
§4 Plus-operators
119
Let Pz V be a bounded operator. Then II Vx112' II Pz Vx112+ 11 Pi Vx112 < (II Pz VII +, II pz VI12+µ) II x111.
and therefore the plus-operator V is also bounded. Conversely, if V is a bounded operator, then it is obvious that Pz V is also a bounded operator. Now suppose that the plus-operator V admits closures, i.e., if follows from y and x -i B that y = B. We verify that Pz V also admits closure. Let Vx,, Pz Vx - y+ and x - B. If the sequences (P2' V) and are fundamental, then by (4.1) is fundamental. Since the operator V admits closure, 0, and therefore Pz Vx 9, i.e., y+ = 0, and the operator Pz V admits Vx,, closure.
Conversely, suppose Pz V admits closure. Then V has the same property. 0, then Pz Vx,, - Pz y, and the fact that PP V can be closed implies that P2 +y - 9; and then it follows from (4.1) that Pi- Vx 0, i.e., y = 9.
For, if Vx - y and x,,
Corollary 4.7: Under the conditions of Theorem 4.6 let .'2 = H. be a Pontryagin space. Then the following assertions are equivalent:
a) V is a continuous operator; b) the operator V admits closure. The eigenvalence of assertions a) and b) follows directly from Theorem 4.6 when we take into account that a finite-dimensional operator (in our case, Pz V) admits closure if and only if it is bounded.
Corollary 4.8: Under the conditions of Theorem 4.6 let 1 = H. and Y2 = IIX be Pontryagin spaces, let cl v = .,Y1 and ,? + ( 1) fl Ker V = 0. Then V is a continuous operator. By Theorem 4.6 if suffices to verify that Pz V is a bounded operator. Let JejJ7 be a basis in.,Yz . Then the operator Pz V can be expressed in the form of a sum x
Pz Vx= i pi(x)el,
(4.2)
i=1
are linear functionals defined on 1v. Continuity of P2 V is equivalent to the continuity of the whole set of functionals (,pi};. Therefore if, where (pl, Ip2,
under the conditions of the Corollary, the operator Pz V were unbounded, would be unbounded. Suppose for definiteness_that pi is unbounded. As is well-known ([XIII] ) Ker ,pl j 9 v, and since lv = .-W1, so ,Yi. By Lemma 1.9.5 there is at
then at least one of the functionals 01, s02, ....
least one positive K-dimensional subspace 91+ in Ker pl (C -Qv). From (4.2) we
have, on the one hand, dim Pz V2+ < x; but on the other hand, since .y,+ + (1) n Ker V = 0, we have dim Pz VY+ = dim V2+ = dim 2+ = x; so we have a contradiction.
120 2
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
In this paragraph we shall assume, if nothing to the contrary is stated, that
V is a continuous strict plus-operator defined on the whole of a J,-space W .Ye, = i and acting into a J2-space Y'2 = Wz O+ i'z, where, as before, J; = P1 - Pi, and P,± are the ortho-projectors from ; on to JY;± (i = 1, 2).
Definition 4.9: A strict plus-operator V is said to be doubly strict if VC is also a strict plus-operator.
We point out that, in general, V` will not be a plus-operator for every plus-operator V (not even if V is a strict plus-operator) (see Exercise 32 below).
4.10: A strict plus-operator V will be doubly strict if and only if V* is a strict plus operator. It is sufficient to use the Definition 4.9 and formula (1.1). Theorem 4.11: the inequality
If V is a strict plus-operator, then there is a S > 0 such that (4.3)
II Vx112 %611X11,
holds for all x E ,0+ (,W, ). The number
it, (V) (II V112+µ+(V))1/2 + 11 VII
can be taken as S. Let X E .:P + (.Yt', ),
11 x 111 = 1. From the strictness of the plus-operator V
and Corollary 4.5 we conclude that the form [ (V ` V - µ+ (V )I, )x, y], is non-negative in .Y°,, and therefore we obtain from the Cauchy-Bunyakovski inequality for any y E YP, with 11 y I1, = 1 that
I [(V`V-h+(V)I,)x,Y]i1z
II(V`V-µ+(V)Il)xll < (II V112+µ+(V))11 Vx11z (for, if (V` V- µ+ (V)I, )x = 0, this is obvious, and in the opposite case we put
y = J,(V`V-µ+(V)I,)xl 11 (V`V-µ+(V)I,)x11, in the above inequality). On the other hand, II (V`V
-µ+(V)I,)XII' ,> µ+(V)- II
V1111 Vx112
and therefore µ+ (V) - II
VII II Vx112
(11
Vx11z, V112+,U+(V))"211
which is equivalent to the inequality II Vx Ilz
µ+(V) (II V112+µ+(V))12+11 VII
§4 Plus-operators Corollary 4.12:
121
Let V be a strict plus-operator, with J v = 01. Then II Pz Vx 112 > (b1,211 x l11
for any X E .JP+ (.,Yj ).
It suffices to note that, because the vector Vx is non-negative, we have II vx112=IIP2+Vxll2+IIPz vx112<211Pz vx112,
and we then use the inequality (4.3). From Corollary 4.5 and Theorem 4.11 we obtain directly Corollary 4.13:
A continuous strict plus-operator V maps every non-negative
(positive, uniformly positive) lineal ? (C.J"i) homeomorphically on to a .
non-negative (positive, uniformly positive) lineal V-W(C.Y(2).
One of the trivial consequences of Theorem 4.11 is the fact that Ker V is a negative subspace. It turns out that a more exact assertion is also true: 4.14 If V is a strict plus-operator and 9 v = Yep, then the subspace Ker V is uniformly negative.
Since the form [ (V c V- µ+ (V )I, )x, y] is non-negative, we obtain from the Cauchy-Bunyakovski inequality that, when x E Ker V and y = Jlx, [(Vcv-µ+(V)IJ)x,y],12=µ+(V)(x,x)i I
-µ+(V)[x,x]I Il (VcV - t +(V)Ii)ll (x,x)i Hence, if V` V = µ+ (V)I1, then Ker V= (B}, and in the opposite case
- [x, X],
11 VcV_ µ '(
V)111
II X 1I
i
,
which proves our assertion.
Let T be an arbitrary linear operator acting from a Ji-space 'Yj into a J2-space W2 and let Pi -q T C 9r, or, what is equivalent, P1 Jr C 9 T. Then the operator T can be represented in the form of a matrix (4.4)
T= 11 Tii 11;.i= 1,
where
T11=P2TP1 IWi,
T12=P2+ TP1_ I-Wi,
T21 = Pi TPi
T22 = Pi TPi
l
.
,
1 ,-WI
.
We note that in the case when the J,-space .,Yi coincides with the J2-space .W'2, i.e., ., = 1(2, Ji = J2, the matrix (4.4) coincides with the matrix (2.1).
Theorem 4.15: Let V be a strict plus-operator and .c1 v = Y1. Then for all -WE .1I1+ . 1) the deficiency def V2 = dim WZ TPi VP is the same. 1
Let Y E .,tl+ and let K be the angular operator of the subspace Y. By
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
122
formula (4.4) the matrix II Vi; II j j=, corresponds to the operator V. Therefore
VIP= ((V1 1 + V12K)xt +(V21+ V22K)xi I xi E Pi) and
P2 V9= ((V11 + V12K)xi I xi E.Wi ), P2 V!= (V11 + V12K).
i.
(4.5)
By Corollary 4.12 we have II(V11+V12K)xi 112
X1+ +Kxi III
(,6/2IIx; 11 1)
and therefore
Ker(V11+V12K)=(B),
(4.6)
and moreover the norms of the operators (V11 + V12K)-1 (see (3.1)) are uniformly bounded with respect to K E ' (=. i ( i , (4.7) < .2/5. It follows from formula (4.5) that the deficiency of V/ coincides with the V12K)(- 1) II
II (V11 +
co-dimension of the subspace (V11 + V12K).Y1+ in,7t . Let Ko be the angular operator of the subspace Yo E Jt (i'1), and let K be an angular operator such that II K- Ko II < (5/,2) II V12 II . Then, from the from the fact that (Iz + V12(K- Ko)(V11 +
(V11 +
V12Ko)(-1))(V11 + V12K0)-01
and V12Ko)(-1)
II V12(K- Ko)(V11 +
II < 1,
we obtain that the dimensions of the subspaces ((V11 + V12K).4i )1 (1 .W'
and
((V11 + V12Ko)
1)1n
2
coincide, i.e., the deficiencies of the subspaces VMo and V2 coincide. By virtue
of the linear connectivity of the operator ball the validity of the theorem follows.
The theorem just proved justifies the introduction of the following
Definition 4.16: The number 6,(V)=dim . z IP2 VV (9E ../l+ (.'Y, ))
is
called the plus-deficiency of a strict continuous plus-operator V with 1'v=.W1.
We now introduce a criterion for the double strictness of a strict plusoperator. Theorem 4.17:
The following assertions for a strict-operator are equivalent:
a) V is doubly strict; b) the subspace VI'o E -it' (.W2) for at least one go E . //+ (,;V,);
§4 Plus-operators
123
c) VIE ,!/+(J1'2) for any
d) for at least one angular operator
Ko E Yl i the operator (V11 + V12Ko)-' exists and is defined on the whole of .YPz ; e) for any angular operators K E J 1 the operator (V11 + V12K)-' exists and is defined on the whole of W2+.
We notice at once that from (4.5) and (4.6) the implications b) a d) and
c) a e) follow immediately. Further, c) a b) since, by Theorem 4.15, the plus-deficiency S+ ( V) does not depend on the choice of Y E -//+. It is therefore sufficient for us to prove a) d) and c) - a). a)
d). It follows from Proposition 4.10 that V* = I I (V*)ij I I 4 j = I
((V*)ij = Vji, l,) = 1, 2)
is a strict plus-operator whenever V is, and therefore from formula (4.6) written for V* with K = 0, (i.e., 2 = .Y'I) we have Ker Vi, I W2 = (B1
and
, 7vt, =,ivt,.
Comparing these relations with (4.6) we conclude that the operator V,1 is continuously invertible on the whole of .02, i.e., d) holds when KO = 0. It can be verified immediately that in this case (V) II V-1 II2 <1-inf,I1
V21V11xi 112 Il xz
112
and therefore µ+/z (V) V,,11' is a contraction. c)
a). First of all we prove that V' is a plus-operator. If this were not so,
there would be a vector (0 ?6 ) xo E .sA+(,H'2) such that V'xo E .Y'- - (.Y11). Let Q
be the angular operator of any maximal negative subspace containing V`xo.
Then K(= Q*) is the angular operator of a maximal positive subspace Ji-orthogonal
Theorem 1.8.11). to V`xo (see Consequently Vcxo [1] (Pi + K),Y"1+ therefore xo [1] V(Pt + K)it" , i.e., and (.Y"z)) is J2-orthogonal to the maximal positive subspace
V(P; + K).Y°; (see Corollary 4.13)-we have obtained a contradiction; so V` is a plus-operator. We verify that V` is a strict plus-operator. In the opposite case v C ,0 + (.Yt'1), and since Ker V = i V J (see 1.12) and since by 4.14 Ker V is a uniformly negative subspace, it follows from Proposition 1.7.4, Corollary 1.7.17, and Proposition 1.1.25 that y, is a maximal uniformly positive subspace (with an angular operator K' E it' = X + ( z , .W'z) with 2 into K' II < 1). The operator V' acts from ,z =.YPz Wi and has, as is easily verified, the matrix representation i = JYi (
Vc= II(V`)ijlI .j=1, where (V`)ii= V1*1 (i= 1,2), and (V`)ij = - Vji(i j). Since
V`.YYz CJfv,,
so
-V2=K'Vi'1 and
V22 =-K'V2*1.
124
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
Hence, we have
inf{[V`x,V`x],IXEJ°++(Jr2)}>1-IIK 112inf{IIV`xll1 xECO'++( 2)}. [ x, x1 2 2 Ilxll2 To show that the operator V` is strict it suffices to show that the quantity on
the right-hand side is positive. To do this we note that it follows from the strictness of the operator V and from Corollary 4.13 that the angular operator of the subspace V.Wi , which coincides with V21 VT I', has a norm less than 1, and therefore 11 V,*1 ' V2*1 11 is also < 1. Therefore (see Proposition 1.8.7)
11xE.9++(e2)} )inf{II I`xz2+Kxz
inf{II
11211
X2 EX2+ ' KEW }
IIxII
inf{IPA V2iiX21IKx2)I1 1Ix2E W2 , KE Y2 2
inf{II(Vi1 II 2*,K)x2 11, x2E X2+ 1172
1
> o.
' 211 V*-' 112(>i 011 V*-' V21 11j)
Definition 4.18: A continuous (J1, J2)-non-contractive operator V with 1'v = W1 is said to be a (J,, J2)-bi-non-contractive operator if V` is a (J1, J2 )-non-contractive operator.
It follows directly from this definition that a (J1, J2)-bi-non-contractive operator is a doubly strict plus-operator. However, not every doubly strict plus-operator is a (J1, J2)-bi-non-contractive operator; as an example we may take the operator "P;, which is a doubly strict operator in the J,-space .'1. The following theorem describes all (J1, J2)-bi-non-contractive operators. For a continuous (J1, J2)-non-contractive operator V with J'v = W1 to be a (J,, J2)-bi-non-contractive operator it is necessary and
Theorem 4.19:
sufficient that it be a doubly strict plus-operator. Necessity: As already noted, this is trivial. Sufficiency: We consider the space 'Yi _ -Y2 @,W1
(4.9)
and we introduce into it a Hilbert metric and an indefinite J1-metric given respectively by the formulae (<X2, x1),
[(X2,x1), where x;, y; E J; (i = 1, 2).
(4.10) (4.11)
§4 Plus-operators
125
It can be verified directly that the Jr-metric is generated by the operator
Jr = J2 Q+ - J1
(4.12)
and that this operator is a selfadjoint involution, and the canonical decomposition of WP is defined be the formulae .or = .-Wr Q+ .mar ,
r =f Q+ 'WIT .
(4.13)
Moreover, it is easy to see from formula (4.11) that if T is an operator densely defined in .Yl and mapping into .2 then the Jr-orthogonal complement to its graph FT = (< Tx, x) I X E 1T) in Wj coincides with the graph of
the operator T`:
rT1-11 = rT` _ ( I Y E 9r).
(4.14)
It follows from the same formula (4.11) that the graph r v of the operator V is a non-negative subspace. Moreover, it is a maximal non-negative subspace. For if it were not, then by Theorem 1.4.5 there would be a vector (x2, xz) in r which was Ji,-orthogonal to r v, and therefore it would follow from (4.14) that xi = V`x2. But since V is a doubly strict plus-operator, xi = 0. We now
use Proposition 4.14 and obtain xz = 0. Therefore r v E -It' (,Wf ), and so r E -11-(fir) (see (4.14) and Theorem 1.8.11),
i.e., [Y,Y]2- [V`Y, V`YJi 5 0 for all yE.2 Corollary 4.20:
If V is a doubly strict plus-operator with
[Vx,Vx]23µ[x,x]1 forµ>0andallxE
v = ,i and if ,,
then
[V`Y, VCYJ 3µ[Y, Y]2 for all yE.02,
(or (what is the same thing) V ` V - µI1 > 0 implies VV` - µI2 S 0), and therefore
µ+(V`)=µ+(V) and max(0; µ_(V`)) =max(0;µ_(V)). It suffices to note that (1/,µ)V is a (J1, J2)-non-contractive operator and then to use Theorem 4.19. Remark 4.21: We leave the reader to compare Theorems 4.17 and 4.19, and
to compile a set of criteria for a (J1, J2)-non-contractive operator to be a (J,, J2)-bi-non-contractive operator.
3
Here we investigate the interconnection between certain classes of plus-
operators V. As before, we shall assume that V is continuous and that
(v=,1.
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
126
Definition 4.22: A plus-operator V is said to be focusing if there is a constant y > 0 such that [ Vx, Vx]2 > y 11 Vx
11
z
for x E
Definition 4.23: An operator V is said to be uniformly (J,, J2)-expansive if there is a constant S > 0 such that [ Vx, Vx]2 >, [x, x] + S x li ;, i.e. 11
V`V- I, >' 0, and to be uniformly (J,, J2)-bi-expansive if the operator V` has the same property.
Theorem 4.24: alent:
For a plus-operator V the following assertions are equiv-
a) b)
V is collinear with a certain uniformly (J,, J2)-expansive operator; V is a focusing strict plus-operator;
c)
µ+(V)> max[O;µ-(V));
d)
there is a vo > 0 such that V ` V - POI,
0.
a) - b). If V is collinear with a uniformly (Jii, J2)-expansive operator, then V is a strict plus-operator (see Corollary 4.5) and there are numbers a > 0 and 6>0 such that [ Vx, Vx]2 '>a [x, x], + S jj x, j12. Therefore when x E ? + (JY,) we have X112 >
[Vx,
II VxlI2,
VI
i.e., V is a focusing plus-operator. b) c). Let V be a focusing strict plus-operator:
(xE e+(
[VX,Vx]2>, y I t'xlI2 I
i))
If a-( V) < 0, then the implication b) - c) follows from the definition of the strictness of a plus-operator V: µ+ (V) > 0.
Now suppose µ_ (V) > 0. We consider operators
Ve = Sf2 V, where
_ (1 - e)P2 + P2-. For sufficiently small e > 0 and x E taking Definition 4.22 into account, Sf2)
(.Y,) we have,
[ Vex, Vex)2 = [(V- ePi V)x, (V- ePi V)x]2 Vx, Vx]2 + (e2 - 2e) [P2 Vx, P2 Vx]2
(y+e2-2e)II VxIiz '> (y+e2-2c)II VexII2, and therefore Ve is a focusing plus-operator. Since [ Vex, Vex]
[ Vx, Vx]2 we
have µ-(V) <µ-(Ve)
(1 - e)2 [ Vx, Vx]2
e2 - 2e < 0 for sufficiently µ+(VE) < (1 - 02A+ (V) < µ+(V). (since
small
e > 0),
and
therefore
§4 Plus-operators
127
c) - d). Suppose the contrary: µ+ (V) > max (0; µ_ (V)) but nevertheless there is no v > 0 such that V`V- PI, : 0. Let the positive numbers v,, v2, v3 satisfy the following inequalities: µ_ (V) < v1 < P2 < P3 < µ+(V). Then (see Corollary 1.1.36) [ Vx, Vx]2 > v; [x, x] 1 (i = 1, 2, 3). From the premise about V we have that there is a sequence of vectors [xn) C -WI with II xn II = 1 such that [IX,,, VXn]2 - v2[xn, xn]1 = [(V`V- P211)X,,, Xn] I - 0 as n - oo.
Since the operator V*H2 V - v2 J1 is non-negative, we have
(V*J2V-v2J1)xn--'0.
(4.15)
Let (x) be a sub-sequence of the sequence (xn) consisting of semi-definite vectors, for definiteness, let us say of non-negative vectors. Then [(V`V- v211)Xn,Xn71 > [(V`V- v31l)X.,X,]I, and therefore
If
C .?
0.
(4.16)
(V*J2V- vIJ1)x,,-'B.
(4.17)
then similarly we obtain
Comparing the relation (4.15) with the relation (4.16) or (4.17) we obtain
x B-a contradiction. The implication d) > a) follows immediately from the fact that 1/ vo V is a uniformly (Ji, J2)-expansive operator.
C Remark 4.25: In proving the implication c) = d) it was established, essentially, that if V is a strict plus-operator and µ+ (V) > v > max [0; A- (V)J, then there is a S, > 0 such that [ Vx, Vx]2 v [x, x] 1 + S, x 11 ;. Moreover, the converse proposition is true: if [ Vx, Vx] 2 v [x, x+ 1 + x 11 1 for all x E .01 ,
v > 0, then µ+(V) > v > max(0;µ_(V)). For, µ+(v) >, v > µ_(V) by Corollary 1.1.36. We verify that v cannot coincide with µ±(V). Suppose, for example, that µ+ (V) = P. Then, by the definition of the number µ+ (V), there would be a sequence (x) C + ( 1) such that [xn ,xn] 1 = 1 and which i.e., [ Vxn, Vxn]2 - . ( V), [ Vxn, VXn]2 - µ+ (V) [Xn, Xn] 1 -- 0, contradicts the inequality [ VXn, VXn]2 - µ+ ( V) [X,,, xn] 1 , Sv x 1 3 sv [xn, I = 0, > 0. Similarly it can be verified that µ_ (V) # P. Thus for any plus-operator V (including even non-strict ones) 11
11
V, V- v11:0 (max10;µ_(V)) < v <µ+(V)),
(4.18)
and when max10;µ_(V)) ?6 µ+(V)
V`V- vlt
0 (max10;µ_(V)) < v <µ+(V)).
This remark enables us to prove the Proposition.
(4.19)
128
4.26:
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
For a uniformly (J,, J2)-expansive operator V to be a uniformly
(J,, J2)-bi-expansive operator it is necessary and sufficient that V` be a strict plus-operator. The necessity is trivial. The sufficiency follows from the fact that, by
virtue of Remark 4.25, it, (V) > 1 > µ_ (V), and therefore we have, from Corollary 4.20, µ+ ( V`) > 1 > (V`). It remains only to apply Remark 4.25 again. We formulate next a theorem which characterizes the `power' of the set of uniformly (J,, J2)-expansive operators. Theorem 4.27: The set of uniformly (J,, J2)-expansive operators ((J,, J2)bi-expansive operators) is open in the uniform operator topology in the set of
all continuous operators acting from ., into .12, and its closure in this topology coincides with the set of all (J,, J2)-non-contractive operators ((J,, J2)-bi-non-contractive operators). Let V be a uniformly (J,, J2)-expansive operator [Vx, Vx]2 [x, x], + V I I - I V'- V112>0. Then
S11x11i(S>0),andlet V'besuch that 6-211 V'- VII
I
[V'x, V'x]2 = [(V+ V'- V)x, (V+ V'- V)x]2 [Vx,Vx]2-(211 V'- VII 11 VII+11 V'- VII2)11x11 [x,x]I + (s -211 V' - VII II VII + II V' - VII2)I1 x11 2 , i.e., V' is a uniformly (J,, J2)-expansive operator. If, in addition, V were a uniformly (J,, J2)-bi-expansive operator, then by Theorem 4.17 0 E p (V ), and therefore for V' from a sufficiently small neighbourhood of V we have 0 E p (V(, ). It then follows from Theorem 4.17 and Proposition 4.26 that V' is a uniformly (J,, J2)-bi-expansive operator. Since the uniform limit of (J,, J2)-non-contractive operators is a (J,, J2)-
non-contractive operator, the closure of the set of uniformly (J,, J2)expansive operators ((J,, J2)-bi-expansive operators) is embedded in the set of
(J,, J2 )-non-contractive operator (J,, J2)-bi-non-contractive operators). It remains to verify that if V is a (J,, J2)-non-contractive operator ((J,, J2)-binon-contractive operator), then it can be approximated in norm by uniformly (J,, J2)-expansive operators ((J,, J2)-bi-expansive operators). To do this we bring into consideration the operators
IV)=,1+eP; +,1-cP1.
(4.20)
It can be verified immediately that when 0 < e <, 1 the operators IV) are uniformly J,-bi-expansive and when c -, 0 they converge in norm to the unit
operator in,. Since [ VIf')x, VIe'>x]2 Vie')
[Ie')x, If')x] _ [x, x] + e II x 11 i,
are uniformly (J,, J2)-expansive operators ((J,, J2)-bi-expansive
§4 Plus-operators
129
operators) which as c --+ 0 approximate in norm the (J,, J2 )-non-contractive ((J1, J2)-bi-non-contractive) operator V. From Theorems 4.24 and 4.27 follows immediately
The set of focusing strict plus-operators (focusing doubly strict plus-operators) is open in the set of all continuous operators and its closure coincides with set of all strict plus-operators (doubly strict plusCorollary 4.28:
operators). We close this paragraph with the Remark 4.29: Plus-operator and sub-classes of plus-operators acting from the anti-space to . i into the anti-space to k2 will be called (as operators acting from W1 into .W2) minus-operators, and the names of sub-classes will be changed correspondingly (for example, (- Jl, - J2 )-non-contractive operators will be called (J,, J2)-non-expansive). All the propositions given above for plus-operators can be reformulated in a natural way for minusoperators. We leave the reader to do this, and later, in using such propositions, we shall refer back to this Remark 4.29.
4
If V is a continuous plus-operator acting from a Jl-space .i, into a
J2-space 2, with ci v = .1, then the operator V` V acts in 01, and therefore it is proper to ask about the description of its spectrum.
Theorem 4.30: Let V and V` be continuous plus-operators. Then
a(V`V)>0. From (4.18) and Theorem 3.24 it follows that a(V`V) C IR. Let 1, such - a E a(V`V), a > 0. Then there is a sequence C or,, 11 x that
(n-- co).
(4.21)
Without loss of generality we may suppose that the sequences [ [ x,,, x"] 1) [ [ Vx,,, 1)I' converge and have the limits 0, y, i and [ [ V` Vx, V`
and S respectively. From (4.21) we obtain (by mulitplying (V ` V+aI, )x scalarly by J, x,, and taking the limit as n oo) y + a/3 = 0 and (multiplying (V ` V + aI, )x scalarly by J, V` Vx and taking the limit as n -p oo) S + Cry = 0.
The first equality implies -y > 0; for, if R > 0, then y > 0 because V is a plus-operator, and if a < 0, then y > 0 because a > 0 by hypothesis and y = - 0. Now from S + ay = 0 and the fact that V is a plus-operator we conclude that S > 0, i.e., S = -y = 0 and therefore 0 = 0 also. Hence, lim
[(V`V-µ+(V)Il)x,,,x],=y-µ+(V)0=0,
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
130
and therefore (cf. proof of Theorem 3.24) (V c V - µ+ (V )Il )xn -p 0 as n oo. Comparing the last relation with (4.21), we obtain xn - 0-a contradiction. We now suppose that V is a continuous operator acting in a J-space W' with
v=.y'. Theorem 4.31: If V is a uniformly J-expansive operator, then its field of regularity contains the unit circle T = (i 11 i; I = 1). Moreover, V is a uniformly J-bi-expansive operator if and only if T C p(V). Suppose E E 7 and that E is not a point of regular type of the operator V.
Then there is a sequence (x,) C . when n S > 0,
with 11 xn 11 = I such that (V - EI )xn
B
oo. Since V is a uniformly J-expansive operator, we have, for some
((Vxn, Vxn] - (xn, xn]
(xn, (V - SI )xn] + S ((V - U )X., xn]
+ ((V-EI)xn, (V-EI)xn] >s 11 xnlli, and therefore xn - 0 when n oo; a contradiction. Now let V be a uniformly J-bi-expansive operator. From what has just been proved and Theorem 1.16, we conclude that IT C p (V ). Conversely, let V be a uniformly J-expansive operator and let I C p (V ). Since
VV'- I= (V- I)(V` -I)-, (Vcv- 1)(V- I)-'(V` -I), so
V`V- I,>0 Corollary 4.32:
implies W-60.
Let V be a focusing strict plus-operator in a J-space .Y. Then
all E such that max (0, µ _ (V)) < I E < µ+ (V) are points of regular type of the operator V, and they are regular if and only if V is a doubly strict focusing plus-operator.
This assertion follows directly from Theorem 4.24, (4.19), and Theorem 4.31.
Exercises and problems 1
Prove that for every strict plus-operator V there is on c = c( V) > 0 such that for
all plus-operators from an e-neighbourhood (in norm) of the operator V the plus-deficiency is the same (M. Krein and Shmul'yan [2] ). 2
Prove that if
V1, VZ are strict plus-operators and S+ (V,) = S. (V2) (M. Krein and Shmul'yan [2] ).
V, - V2 E v'm, then
§4 Plus-operators
131
3
Verify that if V is a strict plus-operator acting from a J,-space .W'i into a J2-space .W2, then Pz V also is a strict plus-operator. Hint: Use the fact [Pz Vx, Pz Vx]z > [Vx, Vx]z.
4
Prove that if V is a J-non-contractive continuous operator acting in a Krein space .Ye and 1P v =.Ye, then the disc 1 = (X II X II < 1) belongs to the field of regularity of
the operator V,,, D C if and only if V is a J-bi-non-contractive operator (M. Krein and Shmul'yan [21). Hint: Use (4.8) and Theorem 4.17. 5
Prove that the conditions a)-e) in Theorem 4.17 are equivalent to the condition f ) 00ap(V*,). Hint: Use Exercise 4.
6
Let .;V = Y+ O+ .JP- be a J-space, dim Y = dim .-W+, and let V be a semi-unitary
operator mapping .# into .N". Verify that V is a strict, but not a doubly strict, plus-operator. Hint: Use Theorem 4.17. 7
Prove that a continuous (J,, J2)-non-contractive operator V is (J,, Jz)-bi-noncontractive if and only if its graph Fv is a maximal non-negative subspace in the Jf-space ,Yr (4.5) (cf. Rintsner [4] ). Hint: Compare (4.14) with Theorem 1.8.11.
8 A plus-operator V is said to be stable if all operators from a certain neighbourhood of it are plus-operators. Prove that assertions a)-d) of Theorem 4.24 are equivalent to the stability of the plus-operator V (M. Krein and Shmul'yan [5] ). 9
Let V= II V;;II?;=1 be a continuous uniformly (J,, Jz)-expansive operator, with 9 v = .YPj. Prove that the operator will be uniformly (J,, J2)-bi-expansive if and only if 0¢ap(V1*1 ).
Hint. cf. Exercise 5 and Proposition 4.26. 10
Prove that the set of all continuous strict plus-operators acting in a J-space J ' forms a subgroup and that µ+(T, T2) > µ+(TI)A+ (T2) > µ_ (T, T2) (M. Krein and Shmul'yan [5] ).
11
Prove that if T, and T2 are strict plus-operators in a J-space .,Y, then 6+ (Ti T2) = 6+ (T,) + 6+ (T2), where, we recall, 6+ (T) is the plus-deficiency of the
operator T (see Theorem 4.15) (M. Krein and Shmul'yan [2] ). 12
Prove that every strict plus-operator in the space II. is doubly strict (Ginzburg [2)).
Hint: Cf. Exercises 4 and 5, using the fact that II, is finite-dimensional. 13
Let V be a (J,, Jz)-non-contractive operator acting from a Pontryagin space H into a Pontryagin space II, , with x = x', and It v = Ih. Then V is a continuous operator (cf. I. lokhvidov [17]). Hint: Verify that the operator V satisfies the conditions of Corollary 4.8.
14
Give an example showing that the condition x = x' in Exercise 13 is essential (Azizov).
15
Let V be a J-non-contractive operator, let .'+ C .,P+, 20+ = M'+
nY[li,
and
dim S'0 < oo. Then V1'+ = Y'+ implies V!'0 = T°+ (Azizov).
Hint: First verify that f°+ C VP°++, and then use the fact that t'0+
is finite-
dimensional. 16
Let .Yy = .Yr+ (D .%(- be a J-space,
-Y.
being separable infinite-dimensional spaces
132
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric with orthonormalized bases (e, )m. (e; )mm respectively. Verify that the linear operator V defined on the basis as follows
Vei =e+ ,, i=0, ± 1,...i
Ve; =e+i,
i= -1, -2,...i
Veo =B
is a J-bi-non-contractive operator and
Y+=CLin[(e,'+ei )0-
(earn E,u+
V'++2'+,
but nevertheless VYO+ ie 2'0++, where Y0+ = 2+ n 2+11 (cf. Problem 15) (Azizov). 17
Prove that if V is a (Ji, J2)-non-contractive operator with J v = .#, and if [Vxo, Vxo] = [xo, xo] for some xo, then [ Vxo, Vy] = [xo, y] for all y E 9 v, and therefore Vxo E V p and V c Vx0 = xo, i.e., xo E Ker(V ` V - I,) (Azizov).
Hint: Use the fact that the graph r v = ( ( Vx, X) I x E / v) is non-negative in the Jr-metric (4.11) and that the vector ( Vxo, xo) (EI'v) is isotropic in it. 18
If V is a J-bi-contractive operator, then a,(V) fl T = 0 (cf. Corollary 2.17) (E. lokhvidov [1]). Hint: Use the result of Exercise 17 applied to the operator V`.
19
Let V be a J-non-contractive operator acting in a J-space Y with / v = W. Prove that Ker (V - EI) C Ker(V` - 41) when E E T; in particular, if V is a J-bi-noncontractive operator and t E T, then Ker (V - EI) = Ker( V` - EI ), and therefore 4 v- t, = . when E o ap( V) (Azizov). Hint: Use Exercise 17.
20
Let V be a J-bi-non-contractive operator, Yo the neutral subspace, and Vxo = Yo. Prove that the operator V induced by the operator V in the )-space .h° _ YY 1 /2'o is J-bi-non-contractive (Azizov). Hint: Use Exercise 17 and Theorem 1.17.
21
Let V be a J-non-contractive operator with f/ v = . , and let Y C Ax(V) fl Ker (V` V- I) and V2' c .T. Prove that Y [1] Y (V) when kµ ;d I (Azizov).
Hint: As in Exercise 8 on § 1, use induction with respect to the parameter p + q, where p, q are the least non-negative integers for which the equalities
(V- XI)Px=B, (V-µI)Qy=B hold for xE', yEY,,(V). 22
Prove that if for an arbitrary J-non-contractive operator V we have Vx = Xx,
Vy=µy and IXI=IAI=1, kite, then [y,x]=0. Hint: Use the result of Exercise 17. 23
Prove that if V is a W-non-contractive operator ( W-non-expansive operator) and Vxo = kxo, Vx, = kx, + x0 when I X = 1, then x0 [1] Ker( V - XI) (cf. Exercise 11 on §2). Hint: Use the fact that [Vx, Vx] - [x, x] is non-negative (non-positive) on l v.
24
Prove that for a a-non-contractive operator V all the root subspaces 2',,(V) (I X I = 1), with the exception of not more than x of them, are negative eigensubspaces (cf. Corollary 2.28) (Azizov). Hint: Use the results of Exercise 22 and 23.
25 A a-non-contractive operator V in a separable space H. can have no more than a countable set of different eigenvalues on the circle T (Azizov, cf. I. lokhvidov [1] ). Hint: Use the results of Exercises 22 and 24. 26
Let ,Yi (i = 1,2) be Ji-spaces. A continuous operator V.-,W - . Z with s v = 01 is called a B-plus-operator if [x, x], > 0, x ;d 6 - [ Vx, VX] > 0. It is clear that V is a
plus-operator. Prove that in the case when .W' = II, every B-plus-operator is a strict plus-operator, but the converse assertion if false ([XVI] ).
§5 Isometric, semi-unitary, and unitary operators
133
27
Verify that in the first assertion of Exercise 26 the condition .YY1 =1I. is essential. Hint: In a J-space which is not a Pontryagin space consider an orthoprojector on to an improper maximal positive subspace and use Proposition 4.14 and Theorem 1.8.11.
28
It is clear that every uniformly (J,, J2)-expansive operator is a B-plus operator. Construct an example showing that even if YP1 = YP2 = rik the converse of this assertion is false.
29
Suppose that V is a B-plus-operator and V` a plus-operator. Show that V` is also a B-plus-operator. In particular (cf. Exercise 12), when W, =.'Y2 = II,,, for every B-plus-operator V the operator V` is also a B-plus-operator.
Hint: Consider an arbitrary FE .,//+ (.)l"2); prove that the subspace (VF)1' negative, and use Theorem I 1.19 (cf. [XVI] ). 30
is
Prove that for a plus-operator V with V('O + (.Y'1) n iv) n p °wY2) ;e 0 it is always true that 31 v C ?+ (.W'2) (Brodskiy [ 1 ] ).
31
Prove that under the conditions of Theorems 4.6 it follows from
-P+,(, r,) n Ker V = 0, 9 v J Wi and the fact that the operator V (see (4.4)) is bounded and V11Jrf t =,,Y2' that V is bounded (Brodskiy [1], cf. I. Iokhvidov [17]). 32
Verify that in a J-space .# = X+ Q+ Y' with dim .R'+ = co the operator V11
V= 11
0
01 1
1-
I
is a strict plus-operator, where V is a semi-unitary operator with V,. + ;d but V` is not even a plus-operator (I. Iokhvidov [XVII]). , - .02 be a strict minus-operator. Then Ker Pi VP; I .Jt"i = [0), and Let V: when 9 v = Ye, the equality .*p2 VP, _ Ye2 is equivalent to the minus-operator V being doubly strict (cf. Ginzburg [2] ). Hint: In proving the first assertion, use the results of Exercise 3 and the equality (4.3), and in proving the second, use Exercise 7 on §1 and Theorem 4.17. The Remark 4.29 has also to be taken into account. Ye+,
33
34
Let
V. .Y 1 -y .02 be a (J,, J2)-bi-non-expansive operator, let 2';(C J';) be
uniformly positive subspaces, W; be the Gram operators of the subspaces TI-L" and let P; be the J;-orthogonal projectors on to Y1J (i = 1,2). Then P2 VP1 I-V[l" is a (W,, W2)-bi-non-expansive operator (Azizov). Hint: Without loss of generality assume that Y; C . Y, (i = 1, 2). Write Vin matrix form relative to the decompositions .0, _ Y, [+] Y[", 2 = M'2 [+] Y?] and calculate the matrices V*J2V(,< J,), VJ, V* (,< J2).
§5. Isometric, semi-unitary, and unitary operators Definition 5.1: A linear operator U acting from a W,-space lei into a W2-space '2 is said to be
I
1) (W,, W2)-isometric if [Ux, Ux]2 = [x, x], when x E iu; 2) (W,, W2)-semi-unitary if it is (W,, W2)-isometric and 9u=7r,; 3) (W,, W2)-unitary if it is (W,, W2)-semi-unitary and Mu = .)2.
In particular, if W, = I, and W2 = 12, then 5.1 is the definition of (ordinary) isometric, semi-unitary, and unitary operators.
134
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
From Definition 5.1 it can also be seen that an arbitrary operator U mapping
a neutral lineal cu into a neutral lineal 11u can serve as an example of a (WI, W2)-isometric operator. Therefore, in contrast to (Hilbert) isometric operators, (WI, W2)-isometric operators in infinite-dimensional spaces can be
unbounded and can have a non-trivial kernel. It is clear that U is a (WI, W2)-isometric operator (respectively a (J1, J2)-unitary operator) if and
only if it is simultaneously a (WI, W2)-non-contractive operator and a (WI, W2 )-non-expansive operator (respectively a (JI, J2)-bi-non-contractive operator and a (JI, J2)-bi-non-expansive operator), and for such operators the corresponding propositions in §4 hold. 5.2 In order that a linear operator V with an indefinite domain of definition should be collinear with some (WI, W2)-isometric operator U, i.e., V= XU (X * 0), it is necessary and sufficient that V should be a plus-operator and a strict minus-operator (or a minus-operator and a strict plus operator) simultaneously.
The necessity is trivial. Now suppose, for example, that V is a plusoperator and a strict minus-operator. Then (see Theorem 4.3) there are constants a > 0 and 0 > 0 such that [ Vx, Vx]2 > a [x, x] I and - [Vx, Vx]2 > -/3[x, x]I when xE qv, and therefore (a -0)[x, x] I < 0. Since _q v is indefinite (by hypothesis), a = (3 > 0, and U = (1 /f) V is a (WI, W2)-isometric operator. Later we shall use the following simple proposition, which follows immediately from the polarization formula (see Exercise 1 on 1 §1). 5.3 For an operator U to be ( W1, W2)-isometric it is necessary and sufficient that [ Ux, Uy12 = [x, y] I for all w, y E c1u.
Remark 5.4: It follows form Proposition 5.3 that if x, y E 9u, then [ x [1] y] a [ Ux [1] Uy], and therefore a lineal Y(C 9u) is isotropic in CAu if and only if U2 is isotropic in 9?u. The possibility was mentioned above of a (WI, W2)-isometric operator with
a neutral domain of definition being unbounded. It turns out that the last condition is not essential; there are even (JI, J2)-semi-unitary unbounded operators.
Example 5.5: Let e _ O+ 0 .e- be an infinite-dimensional J-space with infinite-dimensional let be a maximal dual pair of semi-definite subspaces of the classes h + respectively (see Exercise 4 on § 1.10), let go = + n 9- with dim Yo = 1, and let 9± _ Yo + 2+, where 2'+ are
definite lineals dense in 9+. Then relative to the scalar product ± [x, y] (x, y E `+) the lineals 2+ are Hilbert spaces (see Definition 1.5.9 and Propositions 1.1.23). Therefore there are isometric operators U+ mapping on to r+. We now define an operator U on elements x = x+ + x-, x` E ./%-, by the formula U(x+ + x-) = U+x+ + U-x-. It is easy to see that .W+
this is a J-semi-unitary operator, and that the image of the subspace V+ is the
§5 Isometric, semi-unitary, and unitary operators
135
lineal '+ (not closed in ,Y). From Proposition 5.2 and Corollary 4.13 we conclude that U is an unbounded operator. In this example a sufficient condition for the operator U to be unbounded turned out to be that U.+ was not closed. It turns out that a condition of this sort is also necessary. Theorem 5.6: Let U be a (J1, J2)-semi-unitary operator. Then the following conditions are equivalent.
a) U is a continuous operator; b) U.W' are subspaces; c) 9?u is a subspace.
a) - b) follows immediately from Proposition 5.2 and Corollary 4.13. b) - c) we verify that UM it are uniformly definite subspaces. To do this it
is sufficient (see Proposition 1.5.6) to prove that the subspace U.W' are complete relative to the intrinsic norm I [ Ux, Ux]z 11,2 (x E . i ). But the latter follows immediately from the facts that .ei are Hilbert spaces with the scalar products ± [x, y] 1, and ± [ Ux, Uy] z in . Wl , and U I . i are isometric
operators. Thus U!Pi are uniformly definite subspaces. It only remains to apply Theorem 1.5.7.
c) - a). Since [ Ux, Uy]z = [x, y] 1 for all x, y E M, and M1 is not degenerate, Ker U= (0). Therefore the operator U-1 exists. It follows from the same relation that the fact that U is a (J1, J2)-semi-unitary operator is equivalent to the inclusion. (5.1) U-' C U` Therefore U-1 admits closure and it is defined on the subspace Mu, i.e., U-1
is a continuous operator. Hence if follows by Banach's theorem that U is continuous. Remark 5.7. Essentially it is proved in the implications a) - b) - c) that U.1± and U.Y( are regular subspaces. Corollary 5.8:
Every (J1, J2)-unitary operator is bounded.
Remark 5.9: If . i are Gi-spaces, i.e. 0 ¢ op(Gi) (i = 1, 2), then the relation (5.1) also holds for (G1, G2)-semi-unitary operators, and therefore if Ru is a subspace, then U-1 and U are bounded operators. In particular, all (G1, G2)unitary operators are bounded, and the condition that Ube a (G1, G2)-unitary operator can be rewritten in the form
U-' = U` (= Gi 1U*G2),
Iu= .01,
u=.(z.
(5.2)
2E Now let ,Yi be Ji-spaces, i = 1, 2, and let U = II Ui; II;; - 1 be the matrix representation of a (J1, J2)-semi-unitary (J1, J2)-bi-non-contractive operator.
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
136
By Remark 4.21 and Theorem 4.17 the operator U11 is continuously invertible on the whole of ez, and U21 U11' (=- r) is the angular operator of the uniformly positive subspace U.Yi , and therefore II IF < 1. We bring into consideration the operator z
U(r) =11 U(r )ii 11 i.i=1 = I l r
(I2 -
r*r) - v2
(I2 -
r* (12+ -
r*r) - 1/2
rr*)- 1/2
(12 - rr*)-1/2
I I
(5.3)
operating in the space .J'2 = M2+ c Mj . Using Formula (5.2) it can be verified immediately that U(F) is a J2-unitary
operator, and it follows from 3.21 after a straightforward verification that U(I') is a uniformly positive operator. By the construction of U(F) we have that V = U(r) -' U is a (J1, J2)-semi-unitary (J,, J2)-bi-non-contractive operator, mapping i into Jez , and therefore (see Remark 5.4) also ,fj into z
So V has the matrix representation V = II Vii II?I=1, where V,,: is unitary, Vzz: I -i is a semi-unitary operator, and V12 = 0,
V21 = 0.
Conversely, let Y+ be an arbitrary maximal uniformly positive subspace in -Y2 and let r be its angular operator. We introduce the J2-unitary operator U(F) in accordance with formula (5.3). Let dim Yi = dim Mz , dim I < dim . z , and V,,: e, M2 be unitary, V22: . I Wz a semi-unitary operator, and V21 = 0, V12 = 0. Then V= Vii 11 i J=1 is simultaneously a semi-unitary operator and a (J1, J2)-semi-unitary operator, and U= U(P) V is a (J,, J2)-semi-unitary (J,, J2)-bi-non-contractive operator, mapping l into the given space T+ C'2. We summarize the above argument:
Theorem 5.10:
A one-to-one correspondence has been established between
all triples of operators (r, VI1, V22), where r: w2 contraction, Vi,:. ; - .02+ is a unitary operator, V22:
z
is a uniform
i - .)"
a semi-
unitary operator, and all (J1, J2)-semi-unitary (J,, J2)-bi-non-contractive operators: (r, VI1, V22)
U= U(r)V,
where U(F) is constructed according to formula (5.3), and V= II Viil1?i=,,
V21 =0, V,2=0;
U- (r; V,,, V22), where r = Uz, UI,', V ,
V22 = Pz U(r) ' UPI I . I .
Corollary 5.11: Under the conditions of Theorems 5.10 the operator U is (J,, J2)-unitary if and only if V22 is a unitary operator. Corollary 5.12: If U= II Uii 1I?i=I is a (J,, J2)-semi-unitary (JI, J2)-bi-noncontractive operator, then U21 E .9 - U,2 E 99.
§5 Isometric, semi-unitary, and unitary operators
137
This follows from the implication
9- 0 rE9.4*r*E.91'.- U21EY.. If U = U(r) V is a J-unitary operator and if r E .91,,, then
Corollary 5.13:
C\T C AM. Since
Iz -(12
rE,St ,
-r*r)-ins
IZ - (Iz - rr*)- 112 E .gym,
and therefore U- V E 9,,. Since V is a unitary operator, it only remains to use Theorem 2.11, taking into account that the spectrum of a unitary operator lies on the unit circle.
Remark 5.14: In §§4 and 5 we have not so far investigated the spectral properties of J-non-contractive and J-isometric operators, since later, in §6, we shall establish, by means of the Cayley-Nayman transformation, a connection
between the classes of operators and the classes of J-dissipative and Jsymmetric operators respectively, and, in so doing, a connection between their spectral properties.
3
In this paragraph we introduce and describe a special class of J-unitary
operators, namely, stable J-unitary operators. This description is obtained as a consequence of a more general result.
Definition 5.15: A J-unitary operator U acting in a J-space W is said to be stable if 11 U" 11
<, c
We notice at once that if follows from Formulae (5.2) and (1.2) that U-"
11 U*" 11
= 11 U" 11,
and therefore for a stable U we also have
U-"H
Definition 5.16: A group W
.
U) is said to be amenable if on the space
B( 41) of bounded complex-valued functions on U there is an invariant mean i.e., a linear continuous function f w having the properties a) f,,,(1) = 1;
b) fy,(,p) 3 0 when p E B(P1) and cp(U) > 0 for arbitrary UE c) fyl(,p(U )) = fl(,,( U)) = f#(w) for any U E ((. It follows at once from properties a) and b) that for real functions 'P E B( X11) the inequalities inf(p(U) I UE R1) < fy,(wp) 5 sup(,p(U) I UE 411
hold.
(5.4)
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
138
Remark 5.17: Examples of amenable groups are (see, e.g., [XII] ):
a) soluble groups and, in particular, commutative groups b) compact groups and, in particular, bounded groups of operators in a finite-dimensional space, etc. Theorem 5.18: An amenable group ?/ = (U) of J-unitary operators is bounded (in norm) if and only if there is a dual pair (Y+, 2-) of maximal uniformly definite subspaces invariant relative to W, i.e., U2+ C Y+ for all UE W.
Let W be a bounded amenable group, II U II < c < oo when U E Qi. We consider a function ,px,y: U - (Ux, Uy) (x, U E , , U E 9!). This function is bounded on W: I ox,y (U) I = I (Ux, Uy) I < c2 II x II II y II, and therefore it goes into the domain of definition of the invariant mean f v,. It is easy to see that the function (x, y) f-vl (SoX,y) _ (x, y) I is a non-negative sesquilinear form on '. From formula (5.4) we obtain 2
II x
11 2
< inf Iwo,(U) I UE I'?/] <, (x, x)1 5 sup((U) I UE Qi)
c2
x11
i.e., (x, y)1 is a scalar product equivalent to the original scalar product in the sense of the norms generated by them. By property c) of the invariant mean we have (Ux, Uy)1= f*,(,py( . U)) = f,,,(' ,) = (x, y) 1, (x, y) 1, i.e. the group Uis a group of unitary operators in the new scalar product (x, y) 1. Since (x, y)1 = (Wx, y), where W is uniformly positive, ( W"I'2UW-112) is a
group of unitary operators (in the original scalar product), and moreover W lc2
G=
UW-1,2
W- lie
comments for any U E W with the bounded operator which is self-adjoint and continuously invertible
JW- 1,2,
(O E p (G)).
Consequently the spectral subspaces 2+, k- of the operator G corresponding to its positive and negative spectrum respectively are invariant relative to UW-1,2,
and therefore the subspaces q± = W 122. are and all U E W. Since 9+ is orthogonal to 0, it is immediately verifiable that 2+ is J-orthogonal to 2-, 91+ O+ and from the continuous invertibility of W- 1/2 we have . = ¶+ [+]2-. Moreover, since 1F are definite subspaces in the G-metric, ' are definite subspaces in the J-metric. From Corollary 1.7.17 it follows that 2+- are the operators W1 "2 invariant relative to
uniformly definite subspaces. Thus the first part of the theorem has been proved. Now suppose that a set Q! _ (U) of J-unitary operators (it need not even be assumed that )?/ is a group) has a general invariant dual pair of maximal
uniformly definite subspaces (Y+, Y-), i.e., U22 ± C Y ± (U E ?!). Then .0 = 91+ [+] 2 . We introduce in W a new scalar product, equivalent to the original one (see Theorem 1.7.19)
(x,y)1 = [x+, y+] - [x-, y-], where x = x+ + x-, y = y+ + y-, x+, y± E
_T±.
§5 Isometric, semi-unitary, and unitary operators
139
Relative to this scalar product every operator U E R% is unitary. Therefore, if ' is bounded. mii x I I - 1 1x1 1 , 5 M II x II, then II U II < M/m, i.e., the set Remark 5.19: In proving Theorem 5.18 it has been established that a bounded amenable group W = (U} of operators U (we did not use the fact that they are J-unitary) is similar to the group of unitary operators W1"2 UW-1i2,
and therefore the spectrum of each of these operators is unitary, i.e., a(U) C T.
Corollary 5.20: A J-unitary operator U has an invariant dual pair of uniformly definite subspaces if and only if it is stable. This follows immediately from Theorem 5.18 taking into account that the
group generated by a J-unitary operator is commutative and therefore (see Remark 5.17) amenable.
Definition 5.21: A continuous operator T acting in a Krein space M is said to
be normally decomposable if its spectrum a(T) can be split into two non-intersecting spectral sets a(T) = al (T) U a2 (T) such that the subspaces PQ,(T).
and P°,(T°, invariant relative to T (see Theorem 2.20), are
uniformly definite.
Definition 5.22: A stable J-unitary operator is said to be strongly stable if all J-unitary operators from a certain neighbourhood of it are stable. The following theorem links the concepts of normal decomposability for J-unitary operators with strong stability. Theorem 5.23: decomposable.
A J-unitary operator stable if and only if it is normally
Suppose a J-unitary operator U is strongly stable. By Theorem 5.18 and Theorem 1.7.19 we may assume without loss of generality that U± and therefore U is simultaneously also a (Hilbert) unitary operator. We verify that a(U I ,o+) fl a(U I -) = 0. We assume the contrary, and let o E a(U I '+) n o r (U I 4e- ), and let (e'', e`"2) be an open arc of the unit circle containing the point to = e`"°: 0 < cp, < coo < IP2 < 2a. Let E, denote the spectral function of the unitary operator U. We define an operator
U".":: U",", I (E":-E") =toII (E",-Ejp,)
,
and
UP, =I
It follows from the construction of U" , that it is simultaneously unitary and J-unitary, that (E,, - E",).W is invariant relative to U", and that, having
140
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
chosen the length of the arc (e 1, a"z) suitably, we can make the operator U,,, 0,
arbitrarily close in norm to the operator U. Since U.W± = .
+-
,
(EIP2 - EwN)9 _ (E1vz - E,,)'J+ G) (EIPz - E10-
and therefore (E,, - E,,,).' is a Krein J-space, where J = J I (Eu,, - E,,) .N'. To complete the proof of this part of the theorem it suffices to show that in any neighbourhood of the J-unitary operator toI we can find unstable J-unitary operators. For example, the operators to U(I' ), where U(I') is defined by the
formula (5.3) and r ;4 0 will be unstable J-unstable J-unitary operators. Indeed, the operator U(F) is uniformly positive, U(IF) ;d 1, and therefore the spectrum of U(F) contains at least one point (0< ) Xo qd 1, and then XoEo E a(to U(F )) and I Xol o I ;e 1. Hence, it follows (see Remark 5.19) that
to U(I') is not a stable operator. By choosing the norm of the operator r sufficiently small, we can ensure that the operators toU(r) are sufficiently close to EoI.
Now let
a
J-unitary operator U be normally decomposable, i.e.,
o(U) = o, Uo2, 01 n0`2 = 0, and the subspaces '+ = P,,,.-,Y and '_ = P,,.( are uniformly positive and uniformly negative respectively. Since Uq'+ = Y+, so (cf. 1.11 with Formula (5.2)) U2+1I = -+11, and therefore [2'+, 2'+1]) is a dual pair (invariant relative to U) of uniformly definite subspaces. By Corollary 5.20 the operator U is stable. Therefore (see Remark (5.19)) its spectrum is unitary, i.e., o(U) C T. Just as in Corollary 3.12 it can be verified that P and
Pa, are J-selfadjoint projectors. Let P, and I'2 be corresponding nonintersecting contours, symmetric relative to the unit-circle, which surround a, and o2. Then (see, e.g., [VI]) for perturbations Vf, sufficiently small in norm, of the operator U the spectrum of the operator V = U + Ve will also be divided into two non-intersecting sets of and oZ surrounding F1 and F2 respectively. Moreover, if V is a J-unitary operator, then Pa; are J-selfadjoint projectors,
sufficiently close in norm (because of the closeness of V and U) to the projectors PQ; (i = 1, 2), and therefore (see Problem 10 on §3) they are, together with the latter, projectors on to uniformly definite subspaces. It follows from Corollary 5.20 that V is a stable operator, and therefore U is a strongly stable operator.
In conclusion we introduce one more class of operators closely connected with J-isometric operators. 4
Definition 5.24: Let . 'i be Ji-spaces, i,= 1, 2. A bounded operator U: W, W2 with 1'u = Y1 is said to be partially (J1, J2 )-isometric if Ker U is a regular subspace, and U I (Ker U) [11 is a (J1, J2)-isometric operator.
It follows directly from Definition 5.24 that continuous (J1, J2)-semiunitary operators, and, in particular, (J1, J2 )-unitary operators, are partially (J,, J2)-isometric.
§5 Isometric, semi-unitary, and unitary operators
141
5.25: If U is a partially (J1, J2) -isometric operator, then the operator U` will
also be partially (J1, J2)-isometric. Since (Ker U) [11 is a regular subspace whenever Ker U is, and since the operator U I (Ker U) [11 is continuous, so by Remark 5.7 WU = ?UI tier U1111 is a regular subspace, and therefore Ker U` (= W,, [l1) is also a regular subspace. Since U I (Ker U)111 is (J1, J2)-isometric, it follows that U` I a,, = (U I (Ker U)[-L])-1 is a (J1, J2)-isometric operator. It remains only to use Definition 5.24.
Corollary 5.26: The operator U` which is (J1, J2)-conjugate to a (J1, J2)-semi-unitary operator U is a partially (J2, J1)-isometric operator. Corollary 5.27: Let U be a partially (Jt, J2)-isometric operator. Then U` U is a J1-orthogonal projector on to (Ker U)111.
Exercises and problems 1
Let V be a (J1, J2)-semi-unitary operator (not necessarily bounded); then V ' is a continuous operator (cf, [III] ). Hint: Since 9 v = Y1, the operator V` is bounded (cf. [XXII] ). It remains to use the inclusion (5.1).
.Yt'2, then U is a
2
Prove that if U is a (J1, J2)-semi-unitary operator with (J1, J2)-unitary operator. Hint: Use the result of Exercise 1.
3
Prove that if V is a (J2, J2)-isometric operator, and 9v is closed, and _lv is a non-degenerate subspace, then V is a continuous operator (cf. I. Iokhvidov [5] ).
4
Give examples of (J1, J2)-isometric, unbounded, densely defined operators U: .wt'1-.7'2 for which U-' = U`. Hint: Use the device set out in Example 3.36.
5
Prove that an arbitrary (J1, J2)-unitary operator U can be factorized into a product U = U1 U2, where U1 is a J2-unitary operator with U1 = 12, and U2 is a (J1, J2)-unitary operator iwth U2,Yt'1 =,02:-
Hint: Put U1 = U(r) J2 and U2 = J2 V, where U(P) and V are the same as in Theorem 5.10. 6
Prove that if U is a (J1, J2)-unitary operator, then I U _ (U*U)'/2 is a J-unitary operator. Derive, based on this, the polar expansion U= V U1 of a (J1, J2)unitary operator and prove that V is a (Jr, J2 )-unitary operator and V.W; = "Y2' Hint: Use Theorem 5.10.
7
Prove that a bounded group of 7r-unitary operators in a finite-dimensional space has an invariant dual pair of definite subspaces (Azizov, Shul'man). Hint: Use Theorem 5.18 taking Remark 5.17b) into account.
8
Give an example of a J-unitary operator with a single invariant dual pair (.W, .Yt'-) which is not strongly stable.
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
142
Hint: Take as U, for example, the operator U such that U I+ = I+ and U 1.0- is the shift operator along the orthonormalized basis (e; )m in .Y-. 9
Let U U; 11 ?;., be the matrix representation of a J-unitary operator relative to the canonical decomposition .P' = W+ O+ -W-, and suppose there are open sets 1, and S12 such that a
'=
U (a(U,, + U12K) I KE Jt' ) C 0,, 52, n 512 = 0,
92=U (a(U22 + QU,2) I Q E 3
) C 02,
'P(U) C 52, U 512.
Prove that if p (U) fl (E: I E l = 1) * 0, then U is a strongly stable J-unitary operator (Langer [ 11 ] ). 10
Let NI = ( U) be a group of J-unitary operators acting in a J-space W, let Y be its invariant subspace, and let 2i0 = 2' n Y(1] . Then U2'0 = I° (UE Y!). Hint: Use Remark 5.4.
11
Let V be a continuous J-semi-unitary operator, let 2'° be a neutral subspace, and V91° = 2'°. Prove that 2V is also invariant relative to V and that the operator V 2' /2'° (see Corollary 1.5.8) is J-semi-unitary. induced by V in the f space Hint: Use Proposition 1.11.
12
Under the conditions of Exercise 11, let V be a J-unitary operator. Prove that then V is a J-unitary operator (cf. [XV] ). Hint: Use Exercise 20 on §4 and Exercise 11.
13
Prove that if V is a continuous J-semi-unitary operator, then the uniform definiteness of a subspace 2' is equivalent to the uniform definiteness of the subspace VIP.
Hint: Use the definition of a J-isometric operator and the continuity of the operators V and V-' (see Exercise 1). 14
The spectrum of a J-unitary operator U is symmetric about the unit circle, i.e.,
(XEa(U)) a (k-'Ea(U)).
Hint: Use the relation (5.2) and Theorem 1.16). 15
Let U be a J-unitary operator in Y, and let a, be its bounded spectral set, a, I > 1. Prove that then a2 = (ai) ' is also a spectral set for U, that the projector Pa,uo, is J-self-adjoint, and therefore d'=PQ,ua, is a projectionally complete space. Moreover, if of (I ai I > 1) and a2'= (a(`)-' is a second pair of such spectral sets for U, and a, n a' , = 0, and 2' = Po; u0J, then 2' [1] Y'. Hint: Use Exercise 14. For the rest the proof is entirely similar to that of Corollary 3.12.
16
§6 1
V be a J-isometric operator with k, µ E o ( V) and 1\µ * 1. (cf, [XIV]). 2',,(V) [1] Hint: cf. Exercises 21 and 22 on §4. Let
Then
The Cayley-Neyman transformation We shall find it convenient to introduce this transformation not merely for
linear operators but also in a more general setting-for linear relations (see Definition 1.2) be a given Hilbert space. We consider the set 2' of all linear relations Let T, i.e., of all possible lineals of the space ' x _V1. Thus T = (<x, y)) is the
§6 The Cayley-Neyman transformation
143
linear set of ordered pairs (x, y) (x, y E .Ye), and linear operations in ,Ye x Yt' (and therefore also in T) are defined in the natural way (component-wise). We recall (see § 1.1) that a linear relation T is the graph PA of some linear operator M. If and only if it follows from (w, y) E T and x = B that y = 0; so A: that T = rA = ((x, Ax) ).1 E VA (CAA C Jr'). Such linear relations T are called single-valued linear relations. The sets
VT= (xE.YP there is an yE.J'P such that (x, y) E T) 1 T = (y E ,-,Y I there is an x E .) such that (x, y) E T) J
(6.1)
'
are called respectively the domain of definition and the range of values of the linear relation T. For all T E Y we also introduce the following definitions and notation: Ker T= Ix E.WI (x,0>ET),
Ind T= (yE.Y°j(0,y)ET), (6.2)
-T=(-1)T, XT=((x,Xy)I (x, y)ET)(XEC), T-t=((Y,x)I<x,Y)ET).
(6.3)
(6.4)
Ker T and Ind T are called respectively (see Definition 1.1) the kernel and the indefiniteness of the linar relation T. It is clear that X T, T -' E Y. We also introduce the identical linear relation I= ((x, x) ) XE.,y and the sum of two linear relations T1, T2 E ?:
T1 +T2=(<x,yi+Y2) I<x,Yk)ETk, k= 1,21.
(6.5)
It is clear that Tt + T2 E Y and 9T,+ T2 = cI T, n cT,. In particular, we consider the linear relation
Tx=T- XI
(TE22,XEC)
(6.6)
and by means of it we introduce for a linear relation T its resolvent set
p(T)_ (XEC
I
Ker Tx = (0), 3rr_
),
(6.7)
the spectrum
u(T) = C\p(T)
(6.8)
and, in particular, the point spectrum ap(T) = (XECIKer T>,;e (0)).
(6.9)
Points X E p(T) are called regular points, but 9? E ap(T) are eigenvalues of the linear relation T. For eigenvalues X the lineal Ker Ta = (x E Y' I < x, Xx) E T) is called the eigenlineal, and its non-zero vectors are called eigenvectors of the linear relation T corresponding to the eigenvalues X)'.
We can introduce a further classification of points X(E C) of the point spectrum op (T) of an arbitrary linear relation T by dividing these points into '
It will be convenient later, when Ind T ;d 101, also to suppose that, by definition, oo E oo(T)
(Ca(T)).
144
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
two sub-classes ap, (T) and u,2(T) (cf. Theorem 1.16)
up,,(T)= (XEap(T)I.9?r, ;6 W),
(6.10)
(6.11) ap,2(T)_(XEap(T)I_JiT The continuous spectrum ac(T) and the residual spectrum ar(T) of a linear relation T are defined by the formulae
ac(T) = (X E a(T)\ap(T) 191r, # 4rA =
),
ar(T)_ (XEa(T)\ap(T)I4T,*W)
(6.12) (6.13)
respectively
From (6.7)-(6.9), and (6.12), (6.13) it now follows that a(T) is the union of four mutually non-intersecting parts:
a(T) = up,,(T) U ap,2(T) U ac(T) U ar(T).
The field of regularity r(T) of a linear relation T is the set
r(T) = IX E C\ap(T) I T, =.;?T,)
(6.14)
It is clear that p(T) C r(T) (since e_,), and that r(T)\p(T) C arw(T). It is easy to convince onself that the definitions introduced of the sets p(T), a (T ), ap,, (T), ap,2 (T ), ac(T) and ar(T) for a linear relation T generalize the
definitions of the corresponding sets for closed linear operators A. When T= I A the two sets of definitions are simply identical.
2
For a linear relation T E Y we now introduce the direct (Kr) and the inverse
Cayley-Neyman transformations defined for a non-real parameter To do this we first introduce the corresponding transformations element-
wise, i.e., for all pairs (x, y), (u, v) E ' x, kt(x, y) _ (y - ('z, y - ('x),
(v-u,(v-('u)
(6.15) (6.16)
O A direct calculation will verify the proposition. 6.1 The .Ye x.
transformations kt and kt ' are mutually inverse bijections of
onto. xM.
Now for any (k;4)
E C and S, T E 2 we put
Kr(T)= (ks(x,y))<x,y)ET, Kt '(T)= (kt '(u,v))Es.
(6.17) (6.18)
6.2 The transformation Kt((' # (") maps ? on to Y bijectively, and Kt carries out the inverse transformation.
§6 The Cayley-Neyman transformation
145
This assertion follows directly from 6.1. 6.3
Let TE 9 and S= Kt(T) ((' ;d ) or, equivalently, T= KF t(S). Then Ker S = Ker Tt, Ker T = Ker (CS - CI),
Ind S = Ker Tr, Ind T = Ker St,
(6.19)
!2T= R,,
(6.21)
'Ws ='RT"
V, = RT"
(6.20)
The formulae (6.19) and (6.20) follow directly from a comparison of the definitions (6.2), (6.3), (6.5) with (6.17). The same applies to (6.21) for the verification of which the definitions (6.1) must be brought in again. We now investigate how the spectrum of a linear relation T(E') and its
components are transformed on transition from the linear relation T to another linear relation S = Kr(T). To do this we introduce, for (' # (", a mapping 4;t: C
C in the extended complex plane t = C U (co 1:
(X - (' )(X - .)-t, X ;4 (', X * O°, 00,
(6.22)
= oo.
1,
The function µ = ct(X) maps
Theorem 6.4:
ap,1(T), ap,2(T), ac (T), ac(T), r(S), and p(T) bijectively on to ap.t(S), ap,2(S), ar(S), a,(S), r(S), and p(S)
respectively.
When X ;4 (', X ;d oo, we have by (6.17), (6.15) and (6.22)
S,,=S-µI _ <X,y)ETl
(y-('X) [(Y - rX,
-
II
(y-XX)) <X,y>ET}
from which it follows that when
)X E C
Ker S, = Ker Ta,
(6.23) ,s = ?1Ta. We note that, taking (6.22) and Proposition 6.3 into account, the relations (6.23) remain valid for all X E C if, for any linear relation T, we put
Ker T. = Ind T by definition. 1 We point that, on this understanding, the Definitions (6.10) and (6.11) make sense also for a = oo E ap(T) (see the Footnote on p. 143), and so the assertions of Theorem 6.4 remain vald in this case. This is also true when X =,r E up(T).
146
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
In accordance with this we shall also suppose (cf. (6.7))' that
X=ooEp(T)* (Yr=de, Ind T= (0]]. All the assertions of the Theorem now follow directly from a comparison of the formulae (6.23) with the definitions (6.7)-(6.14). Corollary 6.5:
The function
(µ -
1)1,
00,
µ
l,µ # 00,
µ=1,
(6.24)
µ = 00
effects for the linear relations S and T = KF ' (S) the mapping inverse to the mapping ct (see Theorem 6.4).
3 We return to the most important case for us when the Hilbert space W generating the set 22 (C ,,Y x W) of linear relations is a W-space, i.e., it is equipped with an indefinite W-metric [x, y] = (Wx, y)(x, y E) (see 1.§6.6).
Definition 6.6: A linear relation T(E 2') is said to be W-non-expansive (respectively, W-non-contractive, W-isometric) in W if for all (x, y) E T we have [y, y] 5 [x, x] (respectively, [y, y] > [x, x], [y, y] = [x, x]). It is clear that Definition 6.6 generalizes the corresponding definitions for linear operators V (see Definitions 4.2 and 5.1). The latter definitions are obtained from Definition 6.6 when T = F v is the graph of the operator V. In the case when 91 is a G-space, i.e. [x, y] = (Gx, y) (0 g up(G) we introduce Definition 6.7. For a TE 2' the linear relation
T'= ((u, v) E
X W [y, u] = [x, v] for all (x, y) E 71
is called its G-adjoint. Remark 6.8: It can be seen from this definition that T`(E2) exists for any linear relation T. However, even in the case when T= GA is the graph of a
linear operator A ('A, WA C W) the linear relation T` is not necessarily a graph (i.e., a single-valued linear relation)-see Exercise 1 below. However, comparison of the Definitions 6.7 and 1.1 shows that, when T= FA and VA = .W', these definitions are equivalent, i.e., T` = FA'
Definition 6.9: A linear relation T is said to be G-symmetric if T C T`, and to be G-selfadjoint if T` = T. We note that the condition T C T` is equivalent We recall that the condition Ind T= LO is equivalent to T being the graph of a linear operator. But then by Banach's theorem, in the case when this operator is closed, the condition VT implies that the operator is bounded, and this gives meaning to the notation, oo E p(T) in this case.
§6 The Cayley-Neyman transformation
147
in an obvious way to the requirement that [y, x] = [x, y] for all (x, y) E T, and in this form we generalize it to an arbitrary W-space, and call such T W-symmetric.
When T= FA with CIA = ,e the Definition 6.9 goes over into the definitions (of the graph) of a G-symmetric and a G-selfadjoint operator A respectively (cf, with the Definitions 3.1 and 3.2 respectively). Returning to the general case of a W-space we introduce
Definition 6.10: A linear relation is said to be W-dissipative if Im [y, x] > 0 for all (x, y) E T. When T = PA, this condition is equivalent to the operator A being W-dissipative (cf. Definition 2.1). We examine how the Cayley-Neyman transformation affects one or other of the properties of a linear relation T given in the formulae (6.6)-(6.9). The identity (6.25) [Y - ("x, Y - N - [Y - x, y - x] = 4 Im (' Im [ y, x] is established by a simple calculation. From it follows immediately the
proposition. 6.11 Let the linear relations S, T(E.') be connected by the mutually inverse transformations (cf. (6.15)-(6.18))
S=Kt(T) =[
(u-v,'u-('v)I [u,v)ES). Then the following assertions are equivalent: a)
T is a W-dissipative (respectively, W-symmetric) linear relation;
b)
S is a W-non-contractive linear relation when Im W-non-expansive when Im relation for any (' f).
> 0 and is
< 0 (respectively, a W-isometric linear
It is sufficient to compare the Definitions 6.6, 6.8 and 6.9 with the formulae (6.25)-(6.26).
Remark 6.12. It
is sometimes useful to consider the so-called Waccumulative linear relations T distinguished by the condition Im [y, x] < 0 for all (x, y) E T, and the corresponding linear operators. It is clear that only the
factor (- 1) makes them different from dissipative linear relations and operators respectively. Proposition 6.11 can easily be reformulated in terms of W-accumulative linear relations; it is clear from (6.25) that in the reformulation the roles of the conditions Im > 0 and Im < 0 are interchanged. We add further that maximal W-accumulative operators, and, in particular,
J-accumulative operators, are defined in the natural way, and the theory of them is entirely analogous to the theory of maximal W-dissipative (and J-dissipative) operators developed in §2.
148
4
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
In this and the following paragraphs, returning to the main purpose of our
book, we shall specialize our examination still further, restricting it to a single-valued linear relations T (E a'), i.e., to graphs of linear operators. Applying the transformation Kr to T we shall see to it that S = Kr(T) is also a
single-valued linear relation, i.e., a graph. To do this it is necessary and sufficient, in view of the first of the formulae (6.26), to ensure that (' up(T) (see (6.6) and (6.9)). For, if T = rA = ((x, Ax)) XE fA then destruction of the single-valuedness of the linear relation S = Kr(T) is equivalent (see (6.26)) to the simultaneous satisfaction for some x E 2A of the conditions y - ('x = 0 and
y- 'x * 0, where y=Ax, i.e., x ;;6 6 and r E ap(A) (=ap(T)).
Thus, when (f ;e ) Eap(A) we have S=Kr(T)=rv, where V= Vr is a linear operator for which
9v= (Ax- NxEVA;
Rv= (Ax- JXJXEVa,
which can be written shortly in the form
V= I+ ((' - (') (A - CI)- 1.
(6.29)
Moreover, it is clear that 10 ap(V). A simple calculation (cf. (6.26)) shows that here
A=(CV- 'I)(V-I)-'
(6.30)
A = ('I+(('- )(V-I)-'.
(6.31)
or equivalently
It is clear that we can also work in the reverse order, starting with a linear operator V for which 1 ¢ op(V), and for any (;e ) specify the operator A by the formula (6.30) or (6.31). It is also clear that 9A= (VY - AYE V,,
QA= (("VY-f Y)YE C6,
('0 up (A),
and that the formula (6.28) (or (6.29)) is the inverse of the transformation (6.30) (or (6.31)). Traditionally it is precisely in this way that the direct and inverse Cayley-Neyman transformations of operators are defined, and we, allowing a certain freedom, will keep for them the notations V= Kr(A),
A=Kt'(V). The reader will without difficulty reformulate for this particular case Proposition 6.11 and Remark 6.12 which remain valid, of course, on passing from the operator A to its Cayley-Neyman transform, as we shall call, for brevity, the operator V= Kr(A), and reversely. This proposition can be developed further in several directions in the case when it is possible to choose the parameter (f 96 ) so that (' E p(A). In this case
it follows at once from (6.29) that V= Kr(A) is a bounded operator with Vv=.W.
§6 The Cayley-Neyman transformation
149
Theorem 6.13: Let A be a maximal closed J-dissipative operator in a Krein space and let ( ;4 ) (' E p(A). Then when Im > 0 (Im (' < 0) the operator V= Kt(A) is a J-bi-non-contractive (J-bi-non-expansive) operator and
(V - I).
=. M.
Conversely, if V is a J-bi-non-contractive (J-bi-non-expansive) operator and
1 o ap(V), then the operator A = Kr'(V) is when Im (' >O (Im ('
= A = .N' and (- A )` is a J-dissipative operator. Further, by
6.11, V is when Im > 0 (Im < 0) a J-non-contractive (J-non-expansive) operator. From (6.29) taking account of 1.6 and 1.9 we have
V` = I + ( - (')(A` - I)-' = Kt(A`), and the first pair of Theorem 6.13 now follows from Remark 6.12.
Conversely, when 10 up(V) we have, by Exercise 18 and 19 on §4, (V- 1) = 0, and therefore the operator A = V) (see (6.31)) is densely defined and (exactly the same as the operator -A ` = Kt ' ( V`)) it is under the conditions of Theorem 6.13 a J-dissipative operator. Moreover.(see (6.31) and (6.29) (' E p(A), and, by Proposition 2.7, A is a closed maximal J-dissipative operator.
Remark 6.14: It is easy to understand that Theorem 6.13 can be reformulated and proved if the term ` J-dissipative' is replaced everywhere by the term 'J-accumulative' (see Remark 6.12) and the conditions Im (' > 0 and Im (' < 0 change places. If A is a J-selfadjoint operator (A= A') and (f ;6) (' E p (A ), then U= Kr (A) is a J-unitary operator and 1 E ap(U).
Corollary 6.15:
Conversely, if U is a J-unitary operator, 10 ap(U), and (' ;d (', then A = Kr ' (U) is a J-selfadjoint operator. The first assertion follows from the fact that a J-selfadjoint operator A is
a maximal J-dissipative and a maximal J-accumulative operator simultaneoulsy and therefore by Theorem 6.13 and Remark 6.14, for any (' ;d ',
U= Kr(A) is a bounded J-bi-non-expansive and a J-bi-non-contractive operator simultaneously, i.e. (see §5.1) it is a J-unitary operator. The converse argument proceeds entirely analogously. It is not difficult to obtain the result contained in Corollary 6.15 directly without using Theorem 6.13 (cf. Exercise 2 below). 6.16 Under the conditions of Theorem 6.13, the operators corresponding to bounded uniformly J-dissipative operators A (GOA = .i) are when Im > 0 (Im f < 0) bounded uniformly J-bi-expansive (uniformly J-bi-contractive) operators V = Kr(A) and conversely.
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
150
Suppose, for example, that Im follows
from
(6.25)
(when
> 0. Since Im [Ax, x] > yA II x I Iz, it y = Ax) and from (6.28) when
g=Ax-('x, Vg = Ax - 'x that II
Z11g11Z
and similarly for V` (taking account of the fact that for bounded A we have Im [ - A`x, x] = Im [Ax, x] for all x E ,;V), i.e., V is a uniformly J-biexpansive operator. In the converse of this assertion it has to be taken into account that for a uniformly J-bi-expansive operator Vwe have T C p(V) (see Theorem 4.31) and, in particular, 1 E p(V); otherwise the arguments are analogous.
Returning to the general theory of Cayley-Neyman transformations for linear operators A and V (( ), we V= Kr(A) and A = continue first of all the study started earlier (see 6.2) of spectral questions 5
connected with these transformations. With this purpose we shall explain the connection between invariant subspaces of the operators A and V, and also that between the root lineals of these operators.
6.17 LetA be a linear operator in
,
let (' ;e )(" O ap (A ), and let V= Kr (A ).
If the subspace Y C 9A, if A9 C 91 and (' E p (A 12), then ? C 9 v,
VYC9.9 and 1 a(V I _fl. Conversely, if I C "v, V9 C 2', and I ¢ a(V I Y U ap(V), then 2' C 9A, A2' C .' and E p (A 12). Since (see (6.29)) V = I + ( -) (A - (I) ', and by hypothesis (' E p(A 12'),
i.e., (A - ('I)2' _ 2' and 2so 2'C cDv and
V2' C Y. Moreover, V 12 = Kr (A 12'), A 12' and V 2' are closed (and bounded) operators, and from the condition (' E p (A 12') it follows (see Theorem 6.6) that 1 E p (V 12). The converse implication is verified similarly by means of formula (6.31) and Corollary 6.5. Corollary 6.18:
LetA and V be the same as in 6.17, r¢ ap (A) (a 1 ¢ ap(V )).
Then X E ap(A) a (v = (X - )(X - (')-' E ap(V)) and the root lineals /J',(V) and 2',,(A) coincide. The first assertion follows directly from Theorem 6.4. Now let x E 2',(A), i.e., there is p E N such that (A - XI)px = B. Then
2 = Lin (x, Ax, ..., APx) is a finite dimensional invariant subspace of the
operator A: 2'C 9A, AY C 2' and a(A I .') = (X ), i.e.,
(' E p (A 12')
(because (' ;x-1 X). Therefore (see 6.17) 2' C C1 v, V2 C 2' and
(VI 2') = Kr (A 12). By Theorem 6.4 the number P= (X - ')(X- (')-' is the only point of the spectrum of the finite-dimensional operator V1 Y, and so xE Y C 2',(VI 2') C 9',(V), i.e., 2),(A) C . JV). The converse is proved in a similar way.
Remark 6.19:
It is clear that in Corollary 6.18 the first assertion can be
§6 The Cayley-Neyman transformation
151
reformulated equivalently thus: (v E ap(V) } « (X =
3)(v
-1)-'
E op(A)}.
Unfortunately, the requirement that E p(A I 2) (respectively, 1 E p ( V I 2?) imposed in Proposition 6.17 turns out to be extremely stringent. Later, in Chapter 3, we shall see that in particular cases the `preservation' of certain invariant subspaces under the Cayley-Neyman transformation can be ensured under less burdensome conditions. In all the discussions in §6.4 and §6.5 it might have appeared that, in contrast to the arbitrary number }' ( (') on which some requirement such as ¢ op(A) or (' E p(A) was imposed, the number 1, of which it was always required that 1 0 ap ( V), played some special role. But in fact this role Remark 6.20:
can be played by any number c with I e I = 1, if c ¢ ap(V). This leads only to an
insignificant modification of the formulae for the mutually inverse Cayley transformations themselves: (6.32)
-f)(V-eI)-',
(6.33)
and also to corresponding modifications in the formulation of the theorems
and propositions 6.13-6.19, and in particular to the formulae for the transformation of the point spectra in Corollary 6.18 and Remark 6.19: IX E ap(A)) a (v = e(X - f)(X - (') ' E ap(V)), (v E op(V)) a (X = (("v - e)(v - e)-' E op(A)}. 6 In conclusion we consider the Cayley-Neyman transformation V= K-(A) ((' Pd ¢ ap(A)) linking the operators A and V acting in a Hilbert space . = M1 Q+ Mz (a typical situation for a Krein space). We are interested in the case when the operators A and V admit matrix representations (cf. (4.4)) A = I I AtiJI , j=1, V= V;i 11 ?i= 1, generated by the corresponding ortho11
projectors P1 and P2 (Pkr = Yk, k = 1, 2) (for such a representation if necessary that, for example, -1 C 9A and (' E p(A)). In this case the relation
V(A - I) = A - I, which follows from (6.28), can be rewritten in matrix form as V11
A 11 - Ii
V12
A 12
A11-}"I1
A21
A21
A22 - ("I2
A22 - I2 From this we have at once the relations Vz,
V22
I
I
A21
V11 (A 11 - I1) + V12A21 = A 11 -x'11
(6.34)
V11 A 12 + V12 (A22 - rI2) = A 12
(6.35)
Vz, (A u + rI1) + V22A21 = A21 V21 A 1z +, V22 (A22 - rI2) = Azz - (I2.
152
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
In particular, in the case important for us later, when (' E p(A22), it follows from (6.35) that (6.36)
V12 = -(VI I - I, )A,2 (A22 - CI)-'
and if additionally (' E p(A,, ), the elimination of the operator V,2 from the system (6.34), (6.35) gives
Vii [A,, - ('I1 - A12(A22 - 3'I2)-'A21) = Ali - ("It = A12(A22 - 3'I2)-'A2i or
)(Aii - CIi)
(V1t - Ii)[Ii - A12(A22 - ('I2)-'A21(Aii - It)
(6.37)
It is precisely in this form that the relations (6.36) and (6.37) will be used later in Chapter 3 (Theorem 1.13).
§Exercises and problems 1
If T is a single-valued linear mapping into .w', then T` is a single-valued linear mapping into . if and only if Jr = .Y-
2
Derive Corollary 6.15 directly from the formulae (6.29) and (6.31).
3
Prove that Corollary 6.15 can be generalized as follows. Let A be a maximal J-dissipative operator and a maximal J-symmetric operator simultaneously, and let
((' *)1' E p(A). Then V = Kr(A) is, when Im
> 0 (Im
< 0), a bounded
J-bi-non-contractive (J-bi-non-expansive) operator and a J-semi-unitary operator simultaneously and 1 ¢ ap(V). The converse proposition is true, and so are the analogues of both the direct and converse propositions when the word ' Jdissipative' is replaced by 'J-accumulative' and the conditions Im r > 0 (Im 1' < 0) are replaced by Im (' < 0 (Im (' > 0). Hint: cf. Theorem 6.13 and Proposition 6.11. 4
Let V be a W-non-contractive operator in .", let a be its bounded spectral set, and
let P. be the corresponding Riesz projector. Then when I a I > 1 (respectively invariant relative to V is non-negative (nonI a I < 1) the sub space 2 = is positive). In particular, for a W-isometric operator the subspace 2 = P neutral (cf. Theorem 2.21). Hint: To the graph of the restriction V 1 21 (for any f 1) apply Proposition 6.11, Corollary 6.5 and formula (6.31), and then Proposition 6.17 and Theorem 2.21. 5
Let the operator V be the same as in Exercise 4, and let A with I A I > 1 (I A I < 1)
be a certain set of its eigenvalues. Then C Lin(2'a(V))XEA is a non-negative (non-positive) subspace. In the case of a W-isometric V both these subspaces are neutral. Hint: Prove this by analogy with Corollary 2.22, basing the proof on the result of Exercise 4. 6
Let V be a bounded A-non-contractive operator with Cl v = n.. Then a( V) fl (X I I X I > I I consists of not more than x (taking algebraic multiplicity into
account) normal eigenvalues (cf. Corollary 2.23) (Brodskiy [1]). Hint: Use the results of Exercise 5 on §1, Exercise 22 on §4, then Remark 6.19, Proposition 6.17 and Corollary 2.23.
§6 The Cayley-Neyman transformation
153
7
Let n be the set of eigenvalues of a J-isometric operator V and let A fl (A *)-'10. Then C Lin(f,,(V))xEA is a neutral subspace (cf. Corollary 3.14) ([XIV]). Hint: Use the results of Exercise 5, and Exercise 16 on §5.
8
Let U be a ir-unitary operator in II.. Then its non-unitary spectrum consists of not
more than 2x (taking multiplicity into account) normal eigenvalues situated symmetrically relative to the unit circle T (cf. Corollary 3.15) ([XIV]). Hint: Use the results of Exercise 5 on §1, Exercise 22 on §4, Remark 6.20, and Corollaries 6.18 and 3.15. 9
Let U be a a-unitory operator; then a,(U) = 0 (cf. Corollary 3.16). Hint: Use the results of Exercise 8 on §4.
10
If V is a ir-non-contractive operator (1) v = II. ), if k E ap( V) and I k
1, then the
root lineal 'r( V) can be represented in the form ',,(V) =.il"a [+].,ff,, where dim .4", < oo, V,AI",, C,;l"x,, ff,, C Ker( V- XI), and fix is a non-degenerate subspace (in particular, it may happen that .,ff), = (0)). If d, (k), d2(X ), ... , d,, (X) are the orders of the elementary divisors of the operator V 1 SPX, then
Z 2 XEao(V),IXI=1 J=1
[d'(k)I 2
J
+IXj>1 Z
x
(Azizov [8]; cf. [XIV]).
Hint: Use the results of Exercise 5 on §1, Remark 6.20, Corollary 6.18 and Theorem 2.26. 11
Prove that a J-symmetric operator A is J-non-negative if and only if the satisfies the V= Kt(A) J-isometric operator (; ;4 (') Re[I'(J- V)x, x] > 0 for all xE Jv (Azizov, L. I. Sukhocheva).
condition
12
Prove that if V is a J-unitary operator and Re [('(I- V)x, x) > 0 for some f E C and for all x E ., then a(V) C T (Azizov, Sukhocheva).
13
Let A and V be the same as in Proposition 6.17, and let .9 be a finite-dimensional
subspace. Then (2 C 9A, AY C ?) a (V C 9, v, VV C i) (cf. [III] ). Hint: Use Corollary 6.18. 14
15
Let V be a W-non-contractive operator, and let Vxo = exo, Vx, = ex, + xo with e I = 1. Then xo is an isotropic vector in Ker( V - eI). Hint: Apply the transformation Kt ` (I' * (') to VI te(VV) and use the results of Exercise I1 on §4. Prove that if A C B, where A, B E 2, then Kr(A) C Kr(B) and Kt ' (A) C KF ' (B) when (' ;d f.
Remarks and bibliographical indications on chapter 2 In §§1,4,5 the exposition is carried out at first in the most general form-for operators acting from one space .W1 into another 2 (T:,W, .,Y2), and only later is it made concrete for the case of operators acting in a single space. This
is the first time, apparently, that this has been done (at any rate, so systematically; cf. the monograph [V] and our survey [IV]) and mainly in the
interests of the theory of extensions of operators (see Chapter 5). In this connection it should be borne in mind when reading these notes and especially the bibliographical indications that in them, with rare exceptions, no account
154
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
is taken of generalizations (as compared with the primary sources) made in
the text to the case of operators T : Y,2. §1.1. Adjoint operators (relative to an indefinite metric) were first considered in the space IIx by Pontryagin [1], in J-spaces by I. Iokhvidov [5], [6] and by Langer [2], in G-spaces by I. Iokhvidov [12] and next by Azizov and 1. Iokhvidov [1]. §1.2. The idea of using the indefinite metric (1.4) applied to graphs of linear operators for discovering the properties of the operators themselves was first put forward and used by Phillips [1]. Later Shmul'yan [4], [5] and others developed it. Our exposition here follows Ritsner's monograph [4]. §1.3. The formula (1.10) was established by Azizov. For all the rest of the material see [III]. § 1.4. The Propositions 1.11 and 1.12 are 'folk-lore'. Sources of Theorem
1.13 can be found already in Pontryagin's article [1]. The formulation and proof given in the text are due to Azizov who used an idea of Langer's in [XVI]. Normal points are studied in [X]. Theorem 1.16 represents a certain development of Langer's results [2]. §1.5. The device presented here of passing to the factor-space ,W/,W° was first applied in [XV]; Theorem 1.17 we find essentially in Langer [9]. §2.1. W-dissipative operators were introduced in the book [VI]. ir-dissipative operators were first introduced in a particular case by Kuzhel' [5], [6]; they were studied in detail independently by Azizov [1], [4], [5], [8], and also by M. Krein and Langer [ 3 ] ; J-dissipative operators by Azizov [ 5 ], [ 8 ] E. Iokhvidov [ 1 ], Azizov and E. Iokhvidov [ 1 ] in which the main theorem
of this paragraph was established. A geoemtrical proof of the well-known Lemma 2.8 (see [VII]) is due to Azizov. §§2.2, 2.3. The Definition 2.10 and Theorem 2.11 are taken from [X]. Corollaries 2.12 and 2.13 are due to Azizov, as is all the material in §2.3 (see Azizov [8] ).
§2.4. Theorem 2.20 has been borrowed from [VI] and [XXII]. The remaining material of this paragraph was obtained by Azizov. §2.5, §2.6. Theorem 2.26 due to Azizov generalizes the corresponding result of Pontryagin for 7r-Hermitian operators (see also [XIV], [XVI]). The other results of these paragraphs were also established by Azizov.
§3.1. In an abstract formulation (but in a different terminology) rHermitian and a-self-adjoint operators were first studied by Pontryagin [1], who mentions that his attention was drawn to them by S. L. Sobolev (see the
remark on Chapter 3 below). After Pontryagin they were studied by 1. Iokhvidov [1], [2] (in the 'finite-dimensional' case cf. Potapov [1] ), and in more detail see [XIV]. J-self-adjoint operators were considered by Ginzburg [2], I. Iokhvidov [6] and later (in great detail) by Langer [1]-[3]. G-symmetric and G-self-adjoint operators are encountered partially even in Langer [2], and they are considered in detail in [III], to which we refer the reader for details. Proposition 3.7 and its Corollaries 3.8, 3.9 are found in an article by Azizov
Remarks and bibliographical indications on Chapter 2
155
and E. Iokhvidov [1]. As regards Corollary 3.12 see I. Iokhvidov [6], Langer
[2]. Remark 3.13 is due to Azizov, Corollaries 3.14 and 3.15 go back to Pontryagin [1] (cf. [XIV]). §3.2. In proving Theorem 3.19 it would have been possible to use Proposition 3.7 and to refer to a well-known `definite' result. However, we decided to bring in what we think is a simple proof, due to Azizov. Lemmas 3.20 and 3.21 used in it have such a tangled pre-history that we are inclined to attribute them to 'folk-lore'. Lemma 3.22 is due to Azizov. Corollary 3.25 is found, essentially, in Potapov [1]. Theorem 3.27 in the case of a continuous operator A is due to Ginzburg [2], and in the form in which it is formulated-to Langer [8]. §3.3. All the examples, except Example 3.33, are due to Azizov. §4.1. Plus-operators Vin a real space II, (9) v = II,, V bounded) were first
considered by M. Krein (see M. Krein and Rutman [1], and later in an arbitrary (complex) II, by Brodskiy [1]). The latter, in contrast to our Definition 4.1, imposed the condition ' v = Ilk (we point out that from this condition v n t,+ + = 0 already follows-see Lemma 1.9.5). The definition and the name 'plus-operator' of a plus-operator V in a general J-space itself were introduced in the articles of M. Krein and Shmul'yan [1], [2]. In contrast to Brodskiy, with these authors 9 v = . and the operator V is bounded a priori. For such operators they formulated Theorem 4.3 and presented for it a proof which remains valid in the more general case (see [XVI] for the previous history of this theorem), and they established a classification (which had appeared earlier in a particular case in Brodskiy's [I] ) of plus-operators into strict and non-strict (this terminology itself is due to them).
J-non-contractive (J-non-expansive) operators in a Krein space were intro-
duced and studied by Ginzburg [1], [2], generalizing the corresponding considerations in Potapov's [1] for finite-dimensional spaces. Earlier M. Krein [4] (see also [XIV]) had considered in IIk the so-called non-decreasing linear
operators V: L4 v= Ilk, V bounded and [Vx, Vx] > [x, x] for XE
.?+. We
point out that this inequality means, even in the general conditions of Definition 4.1, that V is a strict plus-operator with µ+(V) > 1 (cf. [XVI]). For plus-operators V in IIk with v = II, Brodskiy [ 1 ] discovered that they are either finite-dimensional (and then, possibly, unbounded), or they are continuous (cf. with our Corollary 4.8). As I. Iokhvidov [17] pointed out, the same Brodskiy article contains essentially all the ideas used in the proof of Theorem 4.6 and its corollaries. For these results in a rather fuller and explicit form, see I. Iokhvidov [17]. T. Ya. Azizov pointed out that the requirement that , v n ,Y+ * 0 imposed in these papers can be omitted in Theorem 4.6. Regarding Corollary 4.7 see I. Iokhvidov [17]. A curious generalization of plus-operators V was recently proposed and investigated in [XVI]: [ Vx, Vx] > µ [x, x] for some µ E fR and all x E H. For such V finite µ±(V) again exist and µ_ (V) 5 µ < µ+(V), and a number of facts were established, many of which probably remain true in the more
156
2 Fundamental Classes of Operators in Spaces with an Indefinite Metric
general situation when V:. '1 -.02 (in the spirit of our Definitions 4.1 and 4.2). §4.2. The concept of a doubly strict plus-operator and the basic facts about such operators in J-spaces were established in the articles of M. Krein and
Shmul'yan [2), [3]; there is a later bibliography of this topic in [IV]. The proof of Proposition 4.10 in the text was given by I. Iokhvidov [17] (there the priority of Yu. P. Ginzburg on this question is pointed out). J-bi-non-contractive (J-bi-non-expansive) operators were introduced and studied by Ginzburg [2]. Theorem 4.19 and its Corollary 4.20 are due to M. Krein and Shmul'yan [2], but the proof in the text is Azizov's (cf. Ritsner [4] ). It embodies most of the earlier known criteria for J-non-contractive operators V to be J-bi-non-contractive (see Ginzburg [1], [2], I. Iokhvidov [14], [17],
M. Krein and Shmul'yan [1], [2]; in connection with the rejection of the a priori requirement for V to be continuous, see I. Iokhvidov [ 17]. See Ritzer [4], E. Iokhvidov [8] for the generalization of the concept of a J-bi-non-expansive operator (cf. Shmul'yan [4] ). §4.3 Focusing plus-operators were first examined by Krasnosel'skry and A. Sobolev [1], later by A. Sobolev and Khatsevich [1], [2], and in detail by Khatskevich [6], [7], [10], [11], [15]; uniformly J-expansive operators-in the book [VI]. Theorem 4.24 was established by Azizov [10] (see Azizov and Khoroshavin [1]). Some of its assertions were obtained independently by M. Krein and Shmul'yan [5]. Theorem 4.27 is contained essentially in [VI]. §4.4. Theorem 4.30 in the 'finite-dimensional' case was established by Potapov [1], in the general case-see Ginzburg [2], M. Krein and Shmul'yan [3]; the proof in the text is due to Azizov. Theorem 4.31 has been borrowed from [VI]; Corollary 4.32 is due to A. Sobolev and Khatskevich [1], [2] §4, Exercise 12. In connection with Remark 4.29 a warning must be given
against mechanical transfer of results of the type in Exercise 12 from plus-operators to minus-operators (respectively, from (J1, J2)-non-contractive to (J,, J2)-non-expansive operators); it must be remembered that the roles of
the subspaces .; ands (respectively, of the projectors P; and P,.-), i = 1, 2, are interchanged. §4, Exercise 26. B-plus-operators in space H were introduced by Brodskiy [1] (see [XVI] ), together with the name `B-plus-operator' itself. In conclusion we point out to the reader that much more information about the operators acting in II spaces that are mentioned in §4 can be found in the monograph [XVI].
§5.1. Isometric (in particular, unitary) operators in infinite-dimensional spaces with an indefinite metric were first considered by M. Krein (see M. Krein
and Rutman [1]), I. Iokhvidov [1], [2], I. Iokhvidov and M. Krein [XIV], [XV]. Proposition 5.2, see M. Krein and Shmul'yan [2]. Example 5.5 in the text was given by Azizov; for the other examples, see I. Iokhvidov [12]. Theorem 5.6 is due to I. Iokhvidov [12]. §5.2. In connexion with Theorem 5.10 see [XIV], M. Krein and Shmul'yan [3], and also cf. Azizov [11]. Corollary 5.13 is found in M. Krein [5].
Remarks and bibliographical indications on Chapter 2
157
§5.3. Theorem 5.18 (and its corollaries) in the case of a commutative group is due to Phillips [3]. In the general case of amenable groups we are inclined to attribute it to 'folk-lore', since the proof given in the text in no way corresponds to that given in [VI], for example, for a single operator. This fact
was formally noted by Azizov and Shmul'yan; see also Exercise 7 on this section. For groups in IIj this result was proved by Shmul'yan without the requirement of amenability. Definitions 5.21 and 5.22 were given by M. Krein [XVII]. Theorem 5.23 is
due to M. Krein [XVII] (for details, see [VII), the proof given here was somewhat modified by Azizov. For similar results for stable and strongly stable J-self-adjoint operators, see Langer [11, McEnnis [1].
§5.4. All the results of this paragraph are in M. Krein and Shmul'yan [3]. §5, Exercises. The operators in Exercise 4 are called J-unitary operators by Shul'man. The condition in Exercise 9 was first considered by Masuda [1].
§6. I. Iokhvidov [1] was the first to apply Cayley-Neumann transformations to operators in spaces with an indefinite matric. Then these transformations were considered in detail in [XIV], [III], and were widely applied by many authors. In §§6.1-6.3 we mainly follow Ritsner [4].
INVARIANT SEMIDEFINITE SUBSPACES
3
In this chapter we shall set out results on one of the central problems in the
theory of operators in Krein spaces and, in particular, in Pontryagin spaces-the problem of the existence of maximal semi-definite invariant subspaces for operators and sets of operators acting in these spaces. It will be assumed that the J-non-contractive operators appearing here are bounded and defined on the whole space.
Statement of the problems
§1 1
We have already encountered the concept of an invariant subspace in
Proposition 2.1.11. We now go into it in detail. Definition 1.1: Let T.
-
be an operator densely defined in a Hilbert
Y. We shall say that the subspace 2' is invariant relative to T if Jr FY = 2' and T: 2' - Y. In particular, the subspace 2' = (0) is invariant
space
relative to any operator T; in this case we put by definition p(T j (0)) = C. We note that if Tis an operator defined everywhere in .0, then the condition is always satisfied and moreover Cr -Y) - It would be IT _ possible in Definition 1.1 to require, instead of the condition that clrfl2' be
I:
dense in 2', the inclusion 2' C 9r, but then the set of invariant subspaces would be impoverished. On the other hand, it would be possible to extend this
class by dropping the condition that C4T r) 2' = 2' and leaving only the condition that it fl 2' C .', but then this would lead to the situation, unnatural in our opinion, when any subspace which intersected )T only along the
vector 0 would be an invariant subspace for T. We therefore stay with the Definition 1.1.
In Pontryagin's foundation-laying work [1] it is proved that (in the terminology we have adopted) every 7r-selfadjoint operator A in IIX has a x-dimensional non-negative invariant subspace 2+, which can be chosen so 158
§1 Statement of the problems
159
that Im a(A 12+) >, 0 (Im a(A 12) 5 0). Further development of this result led to the following problems. Problem 1.2: Does every closed J-dissipative (and, in particular, every J-selfadjoint) operator A have an invariant subspace 2+ E _lt+? If it does, is there an .T+ such that Im a(A I Y+) 3 0 (and for a J-selfadjoint A is there also one such that Im a(A I Y+) S 0)?
Problem 1.3: Let d= (A) be the set of maximal J-dissipative operators A with p(A) n C+ ;e 0 whose resolvents commute in pairs, and let Y+ (C.+) be their common invariant subspace, and let p (A 12'+) n C+ ;'6 0 (A Ed).
have a common invariant subspace k+ E /u+ which Does the family contains Y+ and is such that p(A I - +) n C+ ;6 0 (A E d)? We notice at once that, in such a general formulation, Problem 1.2 has a negative answer even for J-selfadjoint J-positive operators (see Theorem 4.1.10 below).
Example 1.4:
Let W1 = Lin (e) (j W; be a Hilbert space, with II a II = 1,
dim .i = oo, and e 1 Wi We define relative to this decomposition a .
completely continuous selfadjoint operator G in el by means of the matrix G22=aG'22, G=IIGGj1I?i=i, where G11=0, G2'2 is a negative completely continuous operator in . I, f E Xi \RGI II f II = 1, and a > 0 is such that II G II < 1, and we introduce in Ml the G-metric [x, y] = (Gx, y). Since r1 is decomposed into the sum of the neutral one-dimensional subspace Lin (e) and the negative AeI, and since it follows from the inclusion f E.I \,G,, that 0 o ap(G), so Ye1 is a G(')-space with x = 1 (see 1.§9.6). Let P be the orthoprojector from
1 onto .01'. Then PG is
a completely continuous G-selfadjoint operator, which has, as is easily verified, not one non-negative eigenvector. In the space
,N' = 1
O+
2,
W2=
(1.1)
1,
we define a J-metric [(x1,X2),(Y1,Y2)1 _ [J(x1,X2),(y1,Y2)] by means of the operator J (see 1.(3.9)) '1=II JuII%i=1,
Jll=G, J12=(I-G2)1/2,
Let Al be the orthoprojector from .
J21=J2,J22=-G.
on to .,Y1 (C.W), and let the linear
A2 I .'2 = I G I -'. Then A 1 = A 1* > 0, operator A2 be such that Ker ..d2 = and A2 = Az > 0. Since the operator Al is bounded, it follows (see 2.Proposi-
tion 1.9 and 2.Corollary 3.8) that A = Al J+ JA2 = A`30. Since W1 C VA,
we can express A in matrix form relative to the decomposition (1.1): A = II Aii II Ii=1 It can be verified immediately that
A11=PG,
A22= -GI GI-'
A12=P(I-G2)1i2+(I-G2)"21GI-1,
A21=0,
3 Invariant Semi-definite Subspaces
160
is well-known that the selfadjointness of G implies iI G I = c, and
It
G. From the matrix representation of the operator A it follows that 0 O ap(A), i.e., A > 0. Since A 21 = 0 and G I G I -1 eRG = eG, so A 'A C 9A. From this the equality VA" = CAA (n = 1, 2, 3, ...) follows. Therefore if X'+ (E,,#' ) were an invariant subspace relative to A, it would also be invariant relative to therefore CAA = W1 Q+
A 2 = 11 (A 2)tj II a=,:
(A2),1 =(PG)2, (A2)21=0,
(A2)22= PGP(I- G2)1,12-(I- G2)1/2G-1,
(A2)22=I2.
Therefore
if x = (x,, x2) E f+ (xi E .wi, i = 1, 2), then also
A2x= <(PG)2x, + PGP(I- G2)1U2x2- (I- G2)1/2 G- 'x2, x2) E Y+. Since
Ker(A2-I)= (<(I- G2)"2x2i -Gx2) 1 x2E
2)= J
2,
Ker(A 2 - I) does not contain maximal non-negative subspaces. Therefore Y+
Ker(A2 - I) also. Since (A2 - I)JA C J,, so (A2 - I)(Y+ fl 9A) C .0, also. From the relations fly -qA = Y+ and Y+ Ker(A 2 - I) we conclude that (A2 - I)(i+ fl gA) * (B). Since (A2 - I)(2+ fl 9A) C Y+ and all the non-negative subspaces in ., are one-dimensional, there is in, a non-negative eigenvector of the operator A and this will also be an eigenvector
for PG; we have obtained a contradiction, because PG has no non-negative eigenvectors.
We note that for the operator A in the above example p(A)=0 and therefore (see 2.Theorem 2.9) there are no maximal uniformly positive subspaces in !2A. At the same time, in Pontryagin's theorem mentioned above, since A = II., the domain !2A does, by I.Lemma9.5, contain such a subspace. It is therefore expedient to introduce the following
Definition 1.5: We shall say that a set of operators of = (A) in a Krein space = + O+ . satisfies condition (L) and we shall write ..el E (L) if, for any
A E V, there is contained in 9A at least one maximal uniformly positive subspace Y+ (in the general case, different _T+ for different A E /); in particular, if z( should consist of a single operator A, then we would write A E (L) and would assume that ,W C Q)A.
If in Problems 1.2 and 1.3 the operators A satisfy condition (L), then naturally, as we shall see later in Theorem 1.13, the non-negative invariant subspaces are to be sought in the lineal CAA itself. In such a formulation the Problems 1.2 and 1.3 have not yet been solved. Later we shall introduce additional conditions under which they have affirmative solutions.
§1 Statement of the problems
161
Another group of problems-for J-non-contractive operators V with 9v=.W-finds its origin in M. Krein's works [4], [5].
2
Problem 1.6: Does every J-non-contractive operator V have an invariant subspace 99+ E AN+? If yes, then is there an Y+ such that I a(V I Y+) I >1 1? Problem 1.7: Let W = { U) be a commutative family of J-non-contractive operators, and let Y+ be their common invariant non-negative subspace (in brief,
Q/
+ C'+. Is there an
§e+ Eilf+
such
that Y+ C 2+ and
w2+ C 9+? We shall show that Problem 1.7 is such a general formulation has a negative solution even for a family 4! consisting of a single operator U. Example 1.8: Let ,e = W+ O+ .e- be a J-space, and let [e, )O', [ei )i be orthonormalized bases in W+ and *,- respectively. We define on the basis
{ ei )o U { e,- ) i a J-semi-unitary operator U: U e o = eo + el + el ,
Ue=e,+1
(i = 1, 2, ...).
The subspace 91+ = C Lin [ e; ) i (C ,2+) is uniformly positive and invariant
relative to U; moreover it does not admit extension into a maximal nonnegative subspace invariant relative to U. For, suppose Y+ C 2'+ (E,/#+) and U9+ C 2+. Since P+Jv7+ = .1+ (see 1.Theorem 4.5), there must be in se+
a vector of the form eo + x- with x- ;d 0, x- E W-, and together with it the vector U(eo + x-) = eo + ei + ei + Ux-. Since ei E.'+, so eo + xi + Ux- E Y+ and (eo + x-) - (eo + el + Ux-) E 9+. Since 9?+ is non-negative we have ei = (I - u)x-, which, as is easily seen, is impossible; we have obtained a contradiction. We note (cf. Theorem 2.8 and Remark 2.4 below) that in this example the `corner' U12 (= P+ UP- I --) = 0.
Definition 1.9: A subspace . is said to be completely invariant relative to a family of operators J E [ T) if' C 1r and Tie = 2' for all T E J (in this case we shall also write JSe = 2'). To conclude this paragraph we point out that, if .'+ is assumed to be a completely invariant subspace relative to ?l, then the solutions of the problems 1.6 and 1.7 have not yet been found. Later, in §§2-5, we shall indicate sufficient conditions under which these problems have an affirmative solution.
3
Finally, the third of the main directions is that generated by Phillips's paper
[3].
Problem 1.10: Let .4 = (A I be an algebra of continuous operators acting in the whole of a Hilbert J-space W; let art be closed relative to J-conjugation,
3 Invariant Semi-definite Subspaces
162
i.e., A E 4 - A` E . I, and let .1 contain I. Under what conditions on W will every dual pair (2+,.'-) which is invariant relative to -d (i.e., C 2', 2'+ C °-, 2+ [1] 2-) admit extension into an invariant maximal dual pair? Is commutativity of got sufficient for this condition?
Problem 1.11: Under what conditions on a group W = ( U) of J-unitary operators will every dual pair (2+, 2-) which is invariant relative to Pl admit extension into a maximal dual pair invariant relative to Pt? Is commutativity of P/ sufficient for this condition? Remark 1.12: In Problem 1.11 the invariance of the subspaces Y± implies their complete invariance, because Pt is a group and therefore it follows from
U2+ C 2+ and U-'2± C 2+ that U2+ = 2±. Moreover, if (2+, .'-) is an invariant dual pair of any J-unitary operator U, then it is easy to see that (C Lin (U-"9?+)o, C Lin (U-"2- )o) is a dual pair which is completely invariant relative to U and which extends (2+, 2- ), and even more precisely, it enters into any extension of (2+, 2-) into a completely invariant dual pair. We point out, regarding their above formulations, that Problems 1.10 and 1.11 have not yet been solved for the case of arbitrary commutative algebras and groups respectively. In §§4 and 5 below we shall give some sufficient conditions for them to be soluble affirmatively (see also 2.Theorem 5.18). We observe also that, under the conditions of Problem 1.10, the algebra .4 coincides with the linear envelope of its J-self-adjoint elements A (A`) with 11 A 11 < 1, and therefore it is contained in the linear envelope of J-unitary operators
U=(I-A2)in+iA, where
(I-A2)"i2= - 1
27r,
ro
]X(I-A2- XI)-1 dx,
the contour F. contains a(I - A 2) within itself and is symmetric about the real axis and lies in the open right-hand half-plane, and Re , X > 0. If now 2' is a maximal semi-definite invariant subspace of the operator U, then UY = 2' (see 2. Remark 4.21), and therefore U-'2'= Y. Consequently
AY=-
(U- U`)2'= 2i (U- U-')2'C Y.
Conversely, if AY C 2', then it follows from the definition of the operator !-A 2) in that (I -A 2)'/22' C 2', and therefore U2' C 2', and, by virtue of 2.Remark 4.21, U2'= Y. Thus, this Remark 1.12 enables us to go into the investigation of Problem 1.11 only, we leave the reader to make the appropriate reformulation for the Problem 1.10 himself.
4
In this paragraph we point out a connection between Problems 1.2, 1.3 and
1.6, 1.7 respectively.
§1 Statement of the problems
163
Theorem 1.13: Let A E (L) be a maximal J-dissipative operator. Then there is a E p(A) fl C+ such that, for 2+ E _tl+, the following conditions are equivalent:
point
a) 2+ C 9A and AY+ C 2+; where UU = Kr(A) = (A - f1)(A - CI)-' b) UY+ C
is the (see 2. §6.4).
9+,
Cagley-Neyman transform of the operator A at the point
Since A E (L), we suppose (see Definition 1.5) that
C f 1A, and with Im > 211 AP+ II belong to p(A) (see 2.Theorem 2.9). Let 2+ C FDA, 2+ E ail+, and AY+ C 27+. Then all the (' specified above 1W+
therefore all
12+, and since A 12+ is belong to the field of regularity of the operator A continuous, they are regular for this operator. Therefore (A-('I)-'Y+_Y+and
Thus, for the implication a) - b) to hold, it is sufficient to take Imp'>2IIAP+II.
with
We shall now prove that among these there are some for which the reverse implication b) - a) holds. Let 27+ E it(+ and Ur2+ C 2+. It is easy to see (see 2.Proposition 6.16) that a) will follow from b) if and only if (Ur - 1)9?+ = Y+ Let K be the angular operator of the subspace 27+. Since (Ur - I )2+ C 2+,
coincidence of these lineals is equivalent to maximal non-negativity of (see l.Theorem 4.5) (Ur - I)2+, or to the equality + = P+ (Ur - I) (P+ + K),Y+, which, in turn, is equivalent to 0 E p(Urji - I+ + Ur12K), where U-,j (i, j= 1, 2) are the blocks of the operator Ur (see 2.(2.1)). Since it follows from 2.Lemma 2.8, 2.Theorem 2.9, and 2.Formula (6.36) that
Ur12= -(Urii-I+)Ai2(A22-(I so
ur11 - I+ + Ur12K= (Ur11 - I+)(I+ - A12(A22 - I )-'K), and therefore it is sufficient for us to verify that 1 E p (ur11) and it is possible to choose with Im > 211 AP+ II so that II A 12(A22 - ')'II < 1. Using
2.Formula(6.37) and supposing without loss of generality that E p(A 11) we have
(Ur11 - I+)(I+ - A12)(A22 - I )-'A21(A11 - I+)-') =
f)(A11 - I+)-'
This equality implies that R,,;,, - r coincides with W+. Since Ur is a J bi-
non-contractive operator (see 2.Theorem 6.13), OEp(Uii),
II U111 II 5 I
(see 2.Remark 4.21), and therefore we have from I+ - Ut i = U11 (Ur - I+) that 9lr - u; = -W+, which implies, as is easily seen, the inclusion 1 E p (U1 i ), and therefore 1 E p (Ur11). with Im (' > 211 AP+ II We now show that there is a
II A12(A22 - I-)
11 < 1.
such that
164
3 Invariant Semi-definite Subspaces
Since the operator A is J-dissipative 0<
Im[A11x+, x+] + Im[A12x-,x+] + Im[A2ix+,x-] + Im[A22x-,x+]
for all x+M+, x- E.)e- fl 'A. In particular, on substituting instead of x the vector x. (() = Im"('(A22 - ("I-)-'x-, a E IR, for all x- E Y_ with II x II = 1 and +_
A12(A22 -
('I-)-1x-
II A12(A22 - ('I-)-'x-II
x
it is assumed that A 12 (A22 - ('I- )x- # B-for, II A12(A22 - rI-)-'x- II = 0 5 q < 1 for any q > 0), we obtain
(here
II A12(A22- ('I-)-'x- I <
Im[f111x+, x+] Im " ('
otherwise,
+ ImIA,,x+. (A» - tI-)-'x-l
+ Im"(' Im IA22IA22 - ('I_)_ IX-, (A22 - ('1 )-'x
]
From 2.Theorem 2.9 and 2.Corollary 2.8 we obtain that II 1/IM (', and when Im (' > II Re (II we have II (A22 II A22 (A22 - tI )-' II I + 2, and therefore II A2 1
All II II A12(A22 - r)-'x - II S IIIm"(' + Im
1 +2
II
+ Im
Since for 0 < a < 1 and sufficiently large Im r we have IIA11II+IIA21 1+ Im (' Im '(' Im
"r
and since x- (II x-1 II = 1) is arbitrary, we obtain
II A12(A22 - ('I-)-l II < I. 1.14 If A E (L), foA =.W, 27 E,41-, and A: 2- - 2-, then T- is an invariant subspace of the operator A. In accordance with Definition 1.1 it is only necessary to verify that
9A fl Y-= 2-. From 1.Corollary 8.16
it
follows that W+ + 2- = M.
Since ye+ C 9A, we have 9)A =,W++ (2- fl GA), and therefore
implies l 2 = 2-.
W
Theorem 1.13 and Proposition 1.14 show that in future we need investigate only Problems 1.6 and 1.7, and the reader himself will be able to formulate the corresponding consequences for maximal J-dissipative operators A E (L) or for families of them ,.d E (L), and so we shall now dwell on them in the main text. In conclusion we mention that, historically, solution of the Problems 1.6 and 1.7 did not always precede the investigation of Problems 1.2 and 1.3; thus,
for example, Pontryagin's theorem mentioned in §1.1 was obtained earlier than Corollary 2.9 set out below in §2, from which we suggest to the reader in Exercise 10 on §2 that he should derive it.
§1 Statement of the problems 5
165
In the preceding paragraphs of this § 1 we decided that we would investigate
the problem of invariant subspaces mainly for J-non-contractive operators and, in particular, for J-semi-unitary and J-unitary operators. For brevity in the further exposition we introduce
Definition 1.15: We shall say that a family W= (UJ of J-non-contractive operators has: a) property -1, + (respectively, 4 _, (D [ll) if it has a common invariant subspace 2'+ E ft+ (respectively, 2' E Jr, 22+ E Jt+, 2+ [1] 91-);
b) Property i+ (respectively,-,) if each of its common completely invariant non-negative (respectively, invariant non-positive, completely invariant non-negative and non-positive) subspaces (including also (0)) admits extension into a maximal semi-definite subspace (of the corresponding sign) invariant relative to q/; c) Property 4 [lJ if each of its invariant dual pairs (2+,-) with
U9?' = 2+ (u E q/) admits extension into a maximal dual pair invariant relative to W. It is clear from Definition 1.15 that if a family W has the properties ( +, c, W, then it also has the properties 4) ±, (D, 'i) respectively, and the property 4) (4)) is equivalent to having properties 4+ and i,_ (c+ and c _) simultaneously. We note also that Definition 1.15 is applicable, in particular, to a family q/ consisting of a single operator U.
To conclude this section we indicate a device which will enable us later to simplify the proof of whether certain J-non-contractive operators (or families L-I and 4' Li I , and to avoid of such operators) have the properties repetitions in the discussions. 6
Definition 1.16:
Let q/ = (U) be a set of J-bi-non-contractive operators. We shall say that a complex y' of properties of the set q/ (the complex may, in particular, consist of just one property) is 4-invariant relative to q! if this same complex of properties is also possesed by every 4/ = (UJ of J-bi-noncontractive operators acting in the factor-space . = 9901l/ 'o (see 2.(1.11)), where 'o (C,-.?O) is an arbitrary subspace completely invariant relative to >>/. It is well-known that (in our terminology) properties of families of
operators, such as the property of being a group, commutativity, etc., are also 4'-invariant. Moreover a proposition such as follows also holds: 1.17 Let -V/ = (U) be a family of J-bi-non-contractive operators. The property that 199 C .i°° and 2' C U91 implies UY C .-°1 is -,D -invariant, and if
W has this property, if 2'+ is a completely invariant subspace relative to /'+ C .:+, and A'0+ = 2'+ r11+1, then U2'°+ = Yo++.
P7/,
3 Invariant Semi-definite Subspaces
166
Let 2'o be an arbitrary neutral subspace completely invariant relative to ,!/, let
.
= 201] /2o, and let '! = (U) be a family of J-bi-non-contractive
operators generated by the family 4/. If SP is a neutral subspace in . if and if C U 2 ' for every CIE 4 / , then 2 = Lin (x E 1 1 2 E 2') is a neutral subspace in
.Y( and 2 C U. = (Ux E Ux I z E 2'). Then by hypothesis U2' is a neutral subspace, which implies that 02 (= U2') is neutral. Now let 2'+ C O+ and U2'+ = 2+. It follows from this that Yo+ C U2'°+ (C 2+ ). By hypothesis U2'°+ is a neutral subspace, and since it contains Yo +, so (see 1 Proposition 1.17 and 1. Definition 1.13) 4/2'° = 9?0
Definition 1.18: A semi-definite subspace is said to be a maximal invariant (respectively a maximal completely invariant) subspace for the family W if it does
not admit non-trivial semi-definite extensions of the same sign in the set of subspaces invariant (respectively completely invariant) relative to W. A maximal invariant dual pair is defined similarly. It follows directly from Zorn's lemma that 1.19 Every semi-definite invariant (respectively, completely invariant) subspace admits extension into a maximal invariant (respectively, completely invariant) semi-definite subspace of the same sign. A similar proposition also holds for invariant dual pairs').
Lemma 1.20: Let (,Yt°r) be the set of J-spaces, and let (Wj,,x = ( U)) be the set of all families of J-bi-non-contractive operators in which is inherent a
certain complex .Y( of -t-invariant properties, including the property (2' C ?°, 2' C U2 - U2' C 3°). Then either every such family has the property 4)+, or among them there is a family (q/) which does not have the property 4)+ and is such that each of its completely invariant non-negative subspaces is positive.
Suppose a family 4/ _ (U) satisfies the condition of the theorem and nevertheless does not have the property 4+. We assume that 2° is a maximal neutral subspace completely invariant relative to W (see definition 1.18). This enables us to pass to the family L/ _ (U) of J-bi-non-expansive operators acting in the f space . = 201120. In accordance with the condition, this family has the complex of properties .Y( and, by virtue of Exercise 21 on 1.§8, like 4/ it does not have the property 4+. We now show that all non-negative subspaces 2+ which are completely invariant relative to 4/ are positive. Let 2°+ be the isotropic part
of such a subspace .+. Then, by Proposition 1.17, .P°+ is a neutral subspace completely invariant relative to every operator U E 4/, and this implies that complete invariance of 2' = Lin (x E R I R E 2°+) relative to U E I. Since We draw the reader's attention to the fact that a maximal invariant semi-definite subspace is not
assumed to be, generally speaking, a maximal semi-definite subspace, and the concepts of a 'maximal invariant dual pair' and an 'invariant maximal dual pair' (cf. 1.DefinitionlO.l) are, in general, not identical.
§2 Invariant subspaces of a J-non-contractive operator
167
it follows from the maximality of 20 as a neutral subspace `1'o c 2 C completely invariant relative to W that 2'o coincides with 2, i.e., -_o+ = (0). Remark 1.21: Corresponding propositions about properties 4,-, 4,1-L1, can be proved in an entirely similar way (see Exercises 4-6).
Exercises and problems Construct a J-positive completely continuous operator A and verify that if the J-space is not a Pontryagin space, then A -1 has no invariant subspaces 2+ E .,ll' which are contained in CIA (Larionov [9] ).
1
2
Let 4V= (U) be a group of J-unitary operators. Prove that for it the properties: a) amenability; b) solubility; c) commutativity; d) UE Q/= U*E 4I; e) uniform boundedness; f) U` = U* = U-1 are 4)-invariant in the sense of Definition 1. lb. Hint: In proving d) first verify that under the conditions of Exercise 2 U* = U.
3
Let it be a family of J-bi-non-contractive operators. Verify that the property ( every
neutral invariant subspace of the family W admits extension into its completely invariant subspace) is 4?-invariant (Azizov). 4
Let Yl be a certain (D-invariant complex of properties inherent in families W/,.W
U)
of J-bi-non-contractive operators and containing the property the 4)-invariance of which was asserted in Exercise 3. Then either every such family has the property 4)_, or among them there is a family (V/) which does not have the property 4)_ and is such that each of its invariant non-positive subspaces is negative (Azizov). 5
Prove that under the conditions of Lemma 1.20 (respectively, of Exercise 4) either each family 4/j,.x there mentioned has the property 4)+ (respectively, $_), or among these families there is a family (a+/) which does not have this property and is such that it has a
maximal completely invariant non-negative subspace .W+ (respectively, a maximal invariant non-positive subspace Y-) which is definite and _T+ !t ..//+ (respectively,
ll.i/l ) (Azizov). 6
If under the conditions of Lemma 1.20 4+/j,,Nis a group of J-unitary operators, then either every such group has the property 4) Ill (respectively 4)1±]), or there is among them a group (-V/) which does not have this property and is such that all its maximal invariant dual pairs are definite (respectively, which has a maximal invariant definite dual pair (.V+, Y-), 99+ 0.,//±) (Azizov). Hint on Exercise 4-6. They are proved similarly to Lemma 1.20.
Invariant subspaces of a J-non-contractive operator
§2
Let.
= . + O+ W- be the canonical decomposition of a Krein space, and let Vi; 11 ?j=1 be the matrix representation (see 2.(2.1)) of a J-non-contractive operator V relative to this decomposition. 1
V= 11
Theorem 2.1: Every uniformly J-expansive operator V has the property 4)+. If V is a uniformly J-bi-expansive operator, then it has the property 4) and it has a
single pair 2+ C .O+ and 2- C .?- of maximal semi-definite invariant subspaces; moreover they are uniformly definite, I a(V I 2+) I µ+ (V) (> 1), and a ( V I 2 )I <, max(0; µ_(V)) (<1).
3 Invariant Semi-definite Subspaces
168
We begin with the assertion about a uniformly J-bi-expansive operator V. By 2.Theorem 4.13 the unit circle is free from the spectrum of the operator V, and therefore its Cayley transform A = K; ' (V) = i (V + I) (V - I)- I will, by virtue of 2.Proposition 6.16, be a bounded uniformly J-dissipative operator and a(A) fl (- oo, oo) = 0 (see 2.Proposition 2.32).
We write a± = o(A) fl C, and let r± be Jordan contours lying in C+- fl p(A) and surrounding a±. By 2.Theorem 2.21 the subspaces Y'+
= P+J( and 99- = P-LW, where
P±_- tat r. (A-XI)-'dX, ,
(2.1)
are respectively non-negative and non-positive invariant subspaces of the operator A. Since P+ + P_ = I (see 2.Theorem 2.20) Y+ + 9?- _ .W
(2.2)
and therefore (see 1.Proposition 1.25) Y± E ill±. From (2.2) and the invariance of 2 relative to A it follows immediately that Y± are also invariant relative to V(= K, (A) = (A + iI) (A - iI) -' ). Since Visa uniformly J-bi-expansive operator, we have: a) An arbitrary vector y E 2+ is, by 2.Theorem 4.17, expressible in the form y = Vx (x E Y+ ), and by 2.Theorem 4.23 there is a S > 0 such that
[y,y] > [x,x]+6llxl12> III IIYIIZ, i.e. Y+ is a uniformly positive subspace; b) For an arbitrary x E 9?- the inequality - [x, x] > - [ Vx, Vx] + 6 II X II 2 > S 11 X II 2 holds (since Vx E : - ), and therefore Y- is a uniformly negative subspace.
Therefore Y` are Hilbert spaces relative to the scalar products ± [x, y] (x, y E Y±), and so in them V I Y+ is an expansive operator and V I 97 is contractive, which implies the inequalities I a( V I p7+) I> 1 and I a( V Y-) 11. But since (see 2.Remark 4.25) any operator X V with 1
µ+(V)
<X<
1
max[O; µ(V))
such that I a(X M Y+) I > 1 and I a(X M Y-) I< 1 will be uniformly inequalities and I a(V I 2+) I > 'µ+ (V) the J-bi-expansive
I a(VIY-) I <, max[O; µ_(V) also follow. Now let the subspaces Y+ (C ?±) be such that V2+ = 2+ and V22_ C Y_. Just as above we can verify that Y+ are uniformly definite, and moreover Ia(VIY+)I>,µ+(V) and Ia(VI'_)I<,max10;A-(V)). From the properties of the Riesz projectors (2.1) (see 2.Theorem 2.20,e)) we obtain Y'± C 91±, and therefore the operator V has the properties 4)±, i.e., the
§2 Invariant subspaces of a J-non-contractive operator
169
property 4 . At the same time we have verified the uniqueness of the maximal semi-definite invariant subspaces of the operator V.
Now let V be an arbitrary bounded uniformly J-expansive operator. We suppose that we have succeeded in extending the J-space .3r into a J-space .Ye = We+ O+
(J = I+ O+ J), and the operator V into a uniformly J-bi-
expansive operator V. Then, by what has been proved above, the operator 17 has a unique maximal uniformly positive invariant subspace 99+ = Lin (z+ + K9+ 19+ E'+ 0 + ) (K E X (M+ 0 W+, tee- ), 11 K 11 < 1), and therefore Y+ (=.r, n W= Lin (x+ + Kx+ I x+ E ..e+)) E Ill+ (M) and V,-W+ C 9?+. Now, if Y+ C .40 +(,W) and V'+ ='+, then V?+ = T+ also.
Hence, 2'+ c k+ n .re = 2+, which proves that the operator V has the property '1+. Therefore to complete the proof it remains for us to construct the extensions mentioned above of the space X and the operator V.
Let V be a uniformly J-expansive operator. Then A = V` V -I is a uniformly J-positive operator (see 2.Theorem 4.24). By 2.Corollary 3.29 the
operator A has uniformly definite invariant subspaces 2i E _lt±, and Y1 [1] Y1, and so without loss of generality (see 1.§7.6) we shall suppose 9?1 = re+- . Since (A, ' C M±) is equivalent to (V ` V WC W± ), we have VM+ [1] V.W-. Since V is a J-non-contractive operator, so (see 2.Corollary 4.13) V.ye+ is a uniformly positive subspace with the angular operator V22 V111 defined on , v,, and 11 V21 VIi 11 < 1. Let I' be the extension of the operator V22 Vi 1' on to the whole of JY+ and let II F II < 1.
We construct a J-unitary operator U(T) by the formula 2.(5.3). Then
V = U(r) W, where W = U-' (F) V is a uniformly J-expansive operator, and WMe+ C ye+ and W.le+ [1] Wye-. Now as -W+ we take one more copy of .-W+ and we extend the J-unitary operator U(r) into a J-unitary operator U(P) in . e = .re+ O+ .ye by putting r D IF, P .;e+ = 0, i.e., U(P) .)e+ = I+. After this it is sufficient now to construct an extension of the operator W into a uniformly operator defined on Let J bi-expansive f iP. [ Wx, Wx] [x, x] = S 11 x 11 2, S > 0. We define the operator W I ie+ by putting Wx+ = Woox+ + W1ox+ + W2ox+, where x+ E)'e+ and
woo = Wii: J°+ -.Ye+; W10 = ((1 + a)I+ + W12 W2)U2U1o:
+_
"y+,
U,o is a partially isometric operator mapping .ye+ on to Ker W11 (C.W+) and Ker U1o = W20 = W22 W*12((1 + CO I+ + W12W1*2 )-1i2U1o: ,y+ _.O-.
It is immediately verifiable that, when 0 < a < S/(l + 11 W22 112), the operator will be uniformly J-bi-expansive. ,
Let T= 11 Ti; 11 ?;-, be the matrix representation of an operator T, defined everywhere, relative to the canonical decomposition -e = W+ O+ .-e- of the 2
3 Invariant Semi-definite Subspaces
170
J-space .. We bring into consideration the functions
Gi(K+)=K+ T,1+K+T12K+- T21- T22K+,
(2.3)
Gr (K-) = K_ T22 + K_ T21K_ - T21 - T,1K_ ,
(2.4)
whose domains of definition are respectively the operator balls X1 (see 1.8.19).
Lemma 2.2: A subspace Y+ E -tf± with angular operator K± is invariant reative to an operator T with )T= W if and only if the operator K+ is a solution of the equation Gf (K±) = 0 respectively.
Let TY+ C Y+ or, what is equivalent, suppose that for every x+ E W+ + which is a solution of the equation
there is a y+ E
T(x+ + K+x+) = y+ + K+y+), which, in its turn, is equivalent to the system TI2K+x+ = y+ T21x+ + T22K+x+ = K+y+}
(2.5)
We now substitute the value of y+ from the first equation of the system (2.5) into the second and, since x+ is arbitrary in ,.+, we obtain that Gi (K+) = 0. Conversely, suppose Gi (K+) = 0. Then we take as the required y+ the vector defined by the first equation of the system (2.5).
The assertion that TY- C £- a Gr (K_) = 0 is proved similarly.
Definition 2.3: We shall say that an operator T satisfies the condition A+ (A-) and we shall write T E A+ (TEA-) if there is an operator K+ E X +, II K+ II < 1 (K_ E .yl-, II K-1 I < 1) such that Gi (K+) (respectively,
Gr (K_) is a completely continuous operator. Remark 2.4: Using 1.Theorem 8.17 we can by a simple calculation satisfy ourselves that T E A+ (respectively, T E A_) if and only if there is a canonical decomposition ,Y = W' [ + ] ,- (respectively, .0 = R' [ + ],,912- ) such that the `corner' Pi TP1 I i (respectively, Pz TPz I ' ), where P;± are the J-orthogonal projectors on to (i = 1, 2) is completely continuous. Therefore the inclusions T E A+ or T E A_ do not depend on the actual decomposi-
tions of the space as might at first sight appear from the Definition 2.3. It follows from Theorem 2.1 that every uniformly J-bi-expansive operator satisfies the conditions A+ and A_. However, even for such operators there is not always a single decomposition for which both the `corners' are completely continuous simultaneously, or, what is equivalent, there is not a Ko EX' with II Ko II < 1 such that Gi (Ko) E .y'., and GT (Ko) E / O simultaneously (see Exercise 2 below). Nevertheless the following proposition holds:
§2 Invariant subspaces of a J-non-con tractive operator
If U is
2.5
a
J-bi-non-contractive J-semi-unitary
U E A_ ).
{ U E A+)
operator,
171
then
In particular, if U is a J-unitary operator, then
{UEA+) (UEA-). Let U21 E .y'.. It follows from 2.Corollary 5.12 that U12 E .Se., and hence
(UE A-). The assertion about J-unitary operators is proved similarly.
3
Let V be a J-bi-non-contractive operator. We introduce the notation
1+(V)=
V-T+=-V+,I a(VI1+)I %:µ+(V)),
(2.6) (2.7)
IVY- =Y, Ia(VIY )I<,max(o;µ (V)). -T (V) For the proof of Theorem 2.8 below we need the following simple proposition: 2.6 Let
+- E
and let K± respectively be their angular operators and
let !± be invariant relative to an operator T with VT= W.
Then
a(T 19+) = a(Tl l + T12K+) and a(T 12-) = a(T22 + R21K- ). This proposition is a corollary of the fact that the projectors P± map V'± homeomorphically on to +- (see 1.Theorem 4.5) and
TIY+=(P+I
2+)-'(T11+T12K+)P+I Y+
and
TI -T- =(PI
mo)-'(T22+T21K_)P-If-.
Moreover, we shall repeatedly use the following lemma which follows from, for example [VII], Theorem 4.5.6.
Let T be a completely continuous operator acting from a Hilbert space e1 into a Hilbert space ,02. If (K6) and ( F6) are bounded generalized sequences of operators consisting of operators acting from .2 into *1 and converging in the weak operator topology to Ko and Fo Lemma 2.7:
respectively (K6 - KO, Fb - FO), then ( TK6) converges in the strong operator topology to TKO
(TK,
(s)
'' TKo) and F6 TK6
Fo TKo.
Theorem 2.8: If a J-non-contractive operator VE A-, then it has the property If V is a J-bi-non-contractive operator, then
a) VEA+=,)±(C`);4 0 and T(V`);4 0; b) VEA- = f-(V`)= E /V I.,l._ =Y+111,x'+E,7+(V)) and VEA+-,q+(V`)=
(A'+E._tf+IA"+='-[1],
Y-E/-(V));
c) VEA-- Lin (X (V) II X I > 1)C fl(Y+I9+E/+ (V)) and VEA+>Lin(X (V)IXI <1)C fl(-T-191-E/-(V)).
Later, in Theorem 3.9, it will be proved by another method that V has the property i+.
3 Invariant Semi-definite Subspaces
172
Ll Let V be a J-non-contractive operator, and let I = (1 + en)P+ + (1 - en)P-, where 0 < c,, < 1. Then V = VI,, is a uniformly J-expansive
operator: [ VInX, VInx] > [Inx, InX] > [X, x] + en II X II
2.
Since Vn11 = (1 + e,,) V,1, the operator Vn will be uniformly J-bi-expansive if
and only if V is a J-bi-non-contractive operator (see 2.Remark 4.21). It follows from Theorem 2.1 that the operator Vn has an invariant subspace Y' E ..//+. Let Kn denote its angular operator. By Lemma 2.2 it satisfies the equation Gj,(K,,) = 0. Since (see 1.Proposition 8.20) the ball X+ is bicompact in the weak operator topology, we can choose from the sequence (K',} (c,, 0 as n - co) a sub-sequence (K,,) which converges in this topology to a certain operator K o E X + as n oo. Since e,, 0 when n co, KaV,,11 = (I + p,,)K,,V11 , Ko Vii, V2122K,1= (1 - e,t) V22
V,121 = (1 + ea) V21 - V21,
K,1-s V22Ko.
Since it follows from Remark 2.4 that we can without loss of generality assume that V12 E .99., we can use Lemma 2.7 and obtain KnV,112K,, = (1 - c,)K,,V12K;, - KO V12Ko.
These relations enable us to conclude that Ko satisfies the equation Gv (Ko) = 0, i.e., by Lemma 2.2 the subspace SP+ = (x+ + Kox+ I x+ E W+1 (Ea/+) is invariant relative to the operator V.
a) Let V be a J-bi-non-contractive operator. We verify that then I a(V I Y+) I > µ+(V). It suffices (taking Theorem 2.1 and 2.Remark 4.25 into account) for us to deal with the case when µ+ (V) = 1. By Proposition 2.6 we have p(V I 2P+) = p(Vo), where Vo = V11 + V1z_Ko. Since (see 2.Remark 4.21) 0 E p(V11) and V_T+ = Y+, so 0 E p(V, 1) n p(Vo). Since (see Exercise 4 on 2.§4, I = ( X II X I < 1) C p(V11) and V12 E .1 ,, it follows from 2.Theorem 2.11 that D C p(Vo). We _ shall prove that Y+ Ei+(V), i.e.,
ao = l n a(Vo) = 0. Let X E p(Vo) n OD. The operators V. =_ 17,111 + V,112K,, converge strongly, by Lemma 2.7, to the operator Vo, and by Theorem 2.1 and Proposition 2.6 we have X E p(V,,). We now verify that supn(II (V,, - XI+)-' II < oo. To do this it is, by a well-known Banach-Steinhaus theorem, sufficient to show that the set ((V,,- XI')-'x+} is bounded for every x+ E +. Let (V,, - XI+)-'x+ = xi We rewrite this equality in the equivalent form
[I+-e,1[(V0-XI +)-'V1zMI Ko)-I+1(170-XI +)
V1V12K,T)
I+)-'V12(Kn- Ko)]2}x,i _ [I+ - (V0- XI+)-'V12(K,1-Ko)](Vo- XI+)-'x+.
(2.8)
Since
II [(Vo-XI+) V12(Kn-Ko)]2II
52II
(Vo-XI+)-`V,z(K,,-Ko)(V0-XI+)
`V12I,
§2 Invariant subspaces of a J-non-contractive operator 0, and V12 E .v'm and by Lemma 2.7 (Vo - )'I+) -' V12 (K,r - KO)
K,1- K,,
173
(S) i 0,
so (see, for example, [XVIII] )
II(Vo- XI+)-'V12(Ka-Ko)(Vo-XI+)-'V2II -0. Sax+, where II T,, 11 - 0, and (S,1) Therefore (2.8) takes the form (I+ + TT)x,i = is a uniformly bounded sequence. Hence for sufficiently large values of >i we obtain that x i = I+ + T a ) ' S,,x+ and I xi) is a bounded sequence, and therefore (V# < Co. Now let Xo E co. Since ao is an isolated point of the spectrum of the operator I
Vo, there is an open disc Do (C D) whose boundary I'o consists of regular points of this operator and is such that Do fl ao = (Xo). Since Fo is compact we conclude that the set ((V,j - XI+ )-' I Xo E Fo) is uniformly bounded with respect to X and n, and therefore the sequence of Riesz projectors
Pa = - 1
2ai converges strongly to the projector
Po= -tai 1
ro
(Va
- XI+)-' dX
r, (V'o- XI+)-' dX
on the root space 2?(Vo) (see 2.Theorem 2.20). But Pa = 0 for all ri, and therefore Po = 0, which implies that 2'x0(Vo) = 10). Thus I a(Vo) I > 1, i.e., f+ (V) ;.d 0. Now let VE A+. We shall suppose that V21 E 91. (see Remark 2.4). We bring into consideration, by 2.Formula (5.3), the operator U(F) with F = V22 Vi1' (E .gym ). The operator U = U-' (F) V is, together with V, J-binon-contractive and, as is easily verified, II U22 II S 1. Therefore U22+F(I+_P*r)-1/2U12+((I-
V22 =
1,r *)_1/2-I-)U22
is the sum of a compression and a completely continuous operator. This enables us, using the same scheme as we used in proving that V E A_ - 9+ (V) # 0, i.e., again starting from Proposition 2.6 (but this time its second part), to prove that V E A+ = f - (V) * 0. We also notice the following obvious implication:
VEA± a V`EA-. Therefore, if V E A_ (respectively, A+ ), then
(2.9)
V`) ;4 0 (respectively,
1+(V) * 0). So assertion a) has been proved. b) Let 2+ E1+ (V). Then A"- = X+ I-'] E ..!!- and V`. IV- C ./I"-. Since I .,l _ -)= (V`)22 = V22 and (V`)21 = - V,Z, so by Proposition 2.26, a(V` u(V2*2 - V2Q), where Q is the angular operator of the subspace .'!"-. From V 2 E .9'm we conclude (see 2. Theorem 2.11) that [ X I I X I > 1) C )5( V` I ,I'- ). Let I Xo I > 1 and V`xo = Xoxo, 0 * xo E ./V-. Since moreover xo E :1'+ (see
2-Exercise 5 on §6), so xo E Poo and therefore xo E 2+ fl .4'-, which implies (see
3 Invariant Semi-definite Subspaces
174
2.Exercise 17 on §4) VV`xo = xo, i.e., Vxo = (l/Xo)xo and 11/X I < 1-we
have obtained a contradiction of the fact that Y+ Eq+(V). Therefore A"- E,J-(V`). Similarly it can be verified that, if /V_ E7_(V`), then 99+ = T- (-L] E '+ (V), and the second implication in b) holds. c) Let X E ap(V) with I X I ie 1 and let xE Y),(V). Then the spectrum of the restriction of the operator V on to the finite-dimensional invariant subspace
Lin ((V - XI)'x(o consists of the one point (X). By 2.Theorem 1.13 we have x [1] A' for all A' E(+ (V`) if I X I < 1 or for all .4' E j- (V`) if i X I > 1. Using the proposition b) which has already been proved we obtain that xE fl (2- 12- E f-(V)) (I X I < 1) or, respectively, that xE fl (Y+ I Y+ E f+(V)( (I X I > 1). Therefore the inclusions indicated in
assertion c) are valid for every 'x(V) and are therefore also valid for their linear envelope.
Corollary 2.9:
If V is a a-non-contractive operator or a 7r-bi-non-expansive
operator, then f±(V) ;4 0 and the inclusions in assertions b) and c) of Theorem 2.8 hold for V In particular, ir-semi-unitary and 7r-unitary operators have these properties.
This follows from the fact that V21 and V12 are finite-dimensional continuous operators and therefore V E A+ fl A_ . Corollary 2.10:
If V(EA+) is a J-semi-unitary J-bi-non-contractive operator,
then f±(V) ?6 0 and assertions b) and c) in Theorem 2.8 hold for it. In particular, J-unitary operators U E A+ fl A- have these properties. It is sufficient to compare Proposition 2.5 with Theorem 2.8.
Remark 2.11: In accordance with 2.Remark 4.29 all the statements of problems and results in §§1 and 2, as also in §§3-5 later, for J-non-contractive and J-bi-non-contractive operators can be reformulated without difficulty in terms of J-non-expansive and J-bi-non-expansive operators, and this we leave the reader to do.
Examples and problems 1
Give an example of a uniformly I-expansive operator which does not have the property & (and even less, the property 4_) (Azizov). ./P+, J4" (O( and in it Hint: Consider a J-space N' _ .,Y® ( ./P-, dim N1 = dim an operator X U where I X I > I and U is a semi-unitary operator mapping .rY into
2
Let.r = .,Y+ Q+ .YP- be a J-space, .,Y+ and ./P- being two copies of one and the same infinite-dimensional Hilbert space. Verify that the operator V = 11 V1 11 zJ= 1, where VII=AN, V12 = (1/2)I, V21=0, V22=,Y/2I, is a uniformly J-bi-expansive
operator. Prove that there is no Ko E .W + with IlKo ll < 1 such that Gi (Ko) E s/' and Gv (Ko*) E Y. simultaneously (Azizov).
§2 Invariant subspaces of a J-non-contractive operator 3
175
Let V be a J-bi-non-contractive operator with V E A_ (A+ ), let co be the spectral set of
the operator V with ao c (X II X I > 1) (respectively, ao C ED), and let PaO be the
corresponding Riesz projector. Prove that then PPO.f c fl (7+ )+ E/+( V) (respectively, P,,Yf c fl (2- I J'- E f - ( V)) (Azizov). Hint: Use 2.Theorem 1.13 and Theorem 2.8.
Let V be a r-non-contractive operator, let I Xo I > 1, and let 2'o be the isotropic
4
part of the lineal 91a0 '(V) (the possibility that 2',,O(V) = J'x 1(V) = J'o = (B)) is not
excluded). Verify that then dim 22o < dim 22x0(V) and that the subspace 2'o + J'xO( V) is non-degenerate (Azizov).
Hint: Use Theorem 2.8, 2.Theorem 1.13, and 2.Exercise 17 on §4. 5
Let V be a r-non-contractive operator, and let 22i, 222 be arbitrary invariant subspaces of it from .,//+. Prove that then, if I X I > 1, we have
dim(221,(VI2',)+2'i,-'(VIY',))=dim(2,(VI222+. x-'(V IX2))=dim 4(V), and if I X I = 1, then dim 22X( V I I',) = dim 2'a (V 12'2) (Azizov). Hint: Use Exercise 4, Theorem 2.8, 2.Theorem 1.13, and 2.Exercise 17 on §4. 6
Let V be a r-semi-unitary operator in 11 let .J'+ E /+ (V) and to (V I d'+) I > 1, and let ap (V) fl a *-, (V 197+) = 0. Then the operator V' has a single x-dimensional positive, and a single x-dimensional neutral, invariant subspace. When x = 1 the operator V` has no other invariant subspaces from //+, and when x > 1 the power
of the set of them is either not greater than 2,-2 or is not less than that of the continuum (Azizov). Hint: Use Exercise 6 and the hint for it. 7
Let II1 =11+ (D rl_ with 11+ = Linle') and let (e, 11' be an orthonormal basis in r1-. Prove that the linear operator V defined on e+ U (e, )i as Ve+ = 2e+ + ei , Ve; = e;+, (i = 1, 2, ...) is bounded and that its closure is a r-semi-unitary operator which satisfies the conditions of Exercise 6 (Azizov).
8
Let A be a bounded uniformly J-dissipative operator with VA =.1y. Prove that the that J'+ are uniformly operator A has a single pair of invariant subspaces 9 E definite and Im a(A 122+) > 0, Im a(A I Y-) < 0 ([VI] ). Hint: Use Theorems 1.13 and 2.1.
9
Let A be a maximal J-dissipative operator, let A E (L), 1/A = ., and let A12 be an A22-completely continuous operator. Prove that the operator A has at least one invariant subspace 22+ E .,//+, 9,+ C c/A and Im a(A 19'+) > 0 (cf. Langer [2], M. Krein [5], Azizov and E. Iokhvidov [1]). Hint: Use Theorems 1.13 and 2.8. Let A be a maximal r-dissipative operator (in particular, a r-self-adjoint operator) in
10
H. with c/A = M. Prove that it has invariant subspaces 1'± E .. //± such that Im a(A 12'+) > 0, Im a(A I Y'-) < 0. Moreover, Lin(95,(A) I Im X > 01 C 99+ and C Lin(A'1,(A) I Im X < 01 C -Y'-1 (Pontryagin [1], Azizov [4], [8], M. Krein and Langer [3]. Hint: Use Theorem 1.13 and Corollary 2.9. 11
Under the conditions of Exercise 10 let OA < r (regarding the symbol OA, see 2.Exercise 6 on §2) and let the right-hand (respectively left-hand) boundary ray of the corresponding angle form with the positive (respectively negative) semi-axis an angle
,p, > 0 (respectively ,p2 > 0). Prove that the operator A then has no points of the spectrum within the angles (- roe, ,p,) and (- r + p1, r - a2) and if, moreover, Ker A is definite, then the operator A has a single pair of invariant subspaces 2'± E. // that J'` are definite and J'+ = Lin (21,(A) Ker A n.#,, I Im X > 01, and r- = (Lin (2',, (A `), Ker A fl .? - I Im it > 0)) `11 (Azizov [8] ). 1
3 Invariant Semi-definite Subspaces
176
Fixed points of linear-fractional transformations and invariant subspaces §3
1
Let Y = W+ O+ .e- be a J-space and V = 11
V;; 11
;- 1
be a J-bi-non-
contractive operator. It follows from 2.Theorem 4.17 that on the operator ball + _ ,X+( +, -) (see 1.Proposition 8.19) the Krein-Shamul'yan linear-
fractional transformation Fv: K - Fv(K) _ (V21 + V22K)(V11 + V12K) ',
(3.1)
or (in equivalent form) Fv (K) V11 + F+V (K) V12K - V21 - V22K = 0
(3.2)
is properly defined. 3.1 Let K be the angular operator of the subspace Y+ E 11+. Then Fv (K) is the angular operator of the subspace V P+ (E_lf+ ), and therefore the function
Fv maps the ball X+ into If x = x+ + Kx+, then
.yf+.
Vx= V11x+ + V21x+ + V12Kx+ + V22Kx+ = (Vii + V12K)x+ + (V21 + V22K)x+ = y+ + (V21 + V22K)(V + V12K)-'y+,
where y+ denotes the vector (V11 + V12K)x+. This proposition and the writing of the function F+V in the implicit form
(3.2) enables us to introduce the concept of a generalized linear fractional
transformation Fv defined by a bounded J-non-contractive operator V (V v = A e) as a mapping of elements of the ball Jy'+ into the set of subsets of this ball:
Fv(K)= (L' E.yl+I L'V11+L'V12K- V21 - V22K=0).
(3.3)
The following proposition is proved in the same way as 3.1 was: 3.2 &+
Let V be a J-non-contractive operator, let K be the angular operator of E. if+, and let L be the angular operator of the subspace V5F+. Then
F+V (K) = X+ (L), where .Yl+ (L) is defined by the formula in 1.8.9.
Definition 3.3: A subset J ( C .YC+ is said to be invariant relative to the generalized linear fractional transformation Flt if F+V (K) fl X ;4 0 for any K E N. Moreover, the restriction F+V I W is understood to mean the mapping Fv I J': K (E.X) -+ F+V (K) fl x. In particular, if J consists of a single operator KO, then KO is called a fixed point of the transformation Fv . From Proposition 3.2 it follows immediately that
3.4 A subspace 2' E If+ with the angular operator Ko is invariant relative to a J-non-contractive operator if and only if Ko E F+V (Ko), i.e., Ko is a fixed
point of Fv .
§3 Fixed points of linear fractional transformations
177
We note that this proposition coincides with Lemma 2.2, if in the latter we put T = V, a J-non-contractive operator. Proposition 3.4 shows another way of seeking the solution of the problem of invariant subspaces. This way consists in investigating when the function Fv has a fixed point. In order to apply this idea we need some topological concepts and results; we introduce the latter without proof. 2
Definition 3.5: Let E be a Hausdorff linear topological space, and Jt' a subset of it. A mapping F which carries points K E X into now-empty convex subsets
F(K) C E is said to be closed if the fact that generalized sequences (K6) and ( Fa) (F6 E F(Kb )) converge to KO and Fo respectively implies that Fo E F(Ko).
Theorem 3.6: (Glicksberg [IX] ). Let iC be a non-empty bi-compact convex subset in a locally convex Hausdorff topological space E, and let F be a
closed mapping of points K E J into non-void convex subsets F(K) C X. Then the function F has at least one fixed point in J', i.e., there is a point Ko EX such that Ko E F(Ko).
It is easy to see that under the conditions of Theorem 3.6 the following proposition holds: 3.7 The set of fixed points of the mapping F is closed In our case the role of E will be played by the space of linear continuous operators acting from one Hilbert space into another, and the role of ..W by the ball yl + of this space or its closed convex subsets. The topology is the weak operator topology. We now pass on to the key result of this section.
Theorem 3.8: Let ,' =+ O+ Y- be a J-space and V = I I Vii I I i;=, a J-non-contractive operator with V12 .99-; let Fv be a generalized linear-
fractional transformation generated by the operator V according to the formula (3.3), and let .X ( C X + ) be a non-empty convex subset, closed in the weak operator topology, which is invariant relative to F+V. Then the mapping
Fv I JY has at least one fixed point. The set of fixed points of the function Fit X is closed in the same topology. We verify that we are under the conditions of Theorem 3.6. Let KE X. Then it follows from a comparison of Proposition 3.2 with 1.Theorem 8.23 that the non-empty set Fv (K) is bi-compact and convex, and the same is true of the non-empty set Fv (K) fl X. Thus the function Fv I .X carries points from it' into its convex non-empty subsets. It remains to verify that Fv I it is a closed mapping. Let (Kb) be a generalized sequence of elements from .yl, and let F6 E Fv (K6) fl it' and K6 - Ko, F6 - Fo. From the Definition (3.3) we then have F6 V1 i +. F6 V12K6 - V21 - V22K6 = 0.
3 Invariant Semi-definite Subspaces
178
2.7 F6 V - Fo V11, V22K6 - V22Ko, by Lemma and F6 V12Ks - Fo V12Ko, we have Fo E Fv (Ko). Since i' is closed, Fo E -W, i.e.,
Since
Fo E Fv (Ko) n W. It now follows from Theorem 3.6 that Fv I i' has at least one fixed point, and from Proposition 3.7 that the set of all such points is closed.
3
As a corollary from Theorem 3.8 we obtain
Theorem 3.9: A J-non-contractive operator V (EA_) has the property
J-bi-non-contractive operator V (EA+) has the property
+; a
_; a J-unitary
operator U (EA+ n A _) has the property 4? Ill.
Let V be a J-non-contractive operator and V E A_ . Without loss of generality, by virtue of Remark 2.4, we assume that V12 E .9'm. If a nonnegative subspace 2'o with the angular operator Ko is completely invariant relative to V, then it follows from Proposition 3.2 that the weakly closed convex set V, (KO) is invariant (see 1.Theorem 8.23) relative to the generalized
linear-fractional transformation Fv. Therefore by Theorem 3.8 the function Fv I .W+ (Ko) has at least one fixed point Ko which, by Proposition 3.2, will be the angular operator of a maximal non-negative invariant subspace !'o of the operator V, and Yo E .SPo.
Now let V be a J-bi-non-contractive operator, VE A+. By Remark 2.4 we can suppose that V21 E Y., and therefore (V`)12 = - Vzl E Y- Consequently we conclude from Theorem 3.8 that the function Fv° has a fixed point in any convex weakly closed subset from X+ which is invariant relative to Fv'. In particular, if 3 o C Velo C Mo, and Qo is the angular operator of the subspace Mo, then it follows from the invariance of 9101) relative to V` (see 2. Proposition 1.11) and from Proposition 3.2 that the weakly closed convex subset X-* (Qo) (see 1.Theorem 8.23) is invariant relative to Ft'. Let 20 = (x+ + Kox+ I x+ E+ ), where KO is a fixed point of the function Fv , I X *(Qo). Then 20 is invariant relative to V` and is J-orthogonal to 91o
(see 1.Proposition 8.22). Therefore x[01] (Elf-) is a subspace invariant relative to V and containing 91o. The last assertion about a J-unitary operator U is proved similarly. Namely, because of Remark 2.4 we can suppose that U12 E Y,. If (2'+, Y_) is a dual pair with U T+ = -T+, and if K+ are the angular operators of the subspaces 91± respectively, then, as a convex weakly closed subset X in .y1 + which is invariant relative to Fu, we consider the non-empty subset .W = J'+ (K+) n .X_* (K_ )
(see 1.Theorem 8.23). By Theorem 3.8 Fu I.X has a fixed point K+, and therefore 9+ = (x+ + K+ x+ I x+ E ,Y+) is a subspace which is invariant relative to U, contains Y+, and is J-orthogonal to Y_. Therefore (9+,. '+1]) is a maximal dual pair which is invariant relative to U and contains (Y+, Y_).
§3 Fixed points of linear fractional transformations
179
Corollary 3.10: Let V be a 7r-non-contractive or a a-bi-non-contractive operator. Then it has the property If V is a a-unitary operator, then it has the property c Ill. Since V12 and V21 are finite-dimensional operators, VE A+ fl A_, and it only remains to apply Theorem 3.9.
Corollary 3.11: Let 11, =11+ 0 II_ be a Pontryagin space and U be a 7r-unitary operator. Then U is stable if and only if all its eigen-subspaces Ker(U- XI) are non-degenerate.
Z Let U be a stable operator. In accordance with 2.Corollary 5.20 we can without loss of generality suppose that U11± = I l+, and therefore
Ker(U- XI) = (Ker(U - xI) fl II+) [ (D] (Ker(U- xI) fl Ii-). This equality implies that Ker(U - XI) are non-degenerate.
Conversely, suppose that all the Ker(U - XI) are non-degenerate. Then X E aa(U) implies I X I = 1 (see Exercise 5 on 2.§6). Therefore (see Exercise 22
on 2.§5) Ker(U- XI) [1] Ker(U -µI) when X pe µ, and so there are precisely p (0 < p < x) different eigenvalues X1, X2, .. X, of the operator U to which correspond non-negative eigenvectors. Since the Ker(U- X,I) (i = 1, 2, . . ., p) are non-degenerate, it follows that
,
Ilk = Ker(U- X,I) [+]Ker(U- X2I) [+]
[+]Ker(U- X I) [+].-I
where
.il'= [Ker(U- X1I) [+]Ker(U- X21) [+] ... [+]Ker(U- X I)J and U,4' C 'U in accordance with 2. Proposition 1.11, and by construction U I A has no non-negative eigenvectors. By Corollary 3.10 /V is a negative subspace and therefore U I N is a unitary operator relative to the scalar product - [x, y] 1,/V. Consequently u is a unitary operator relative to the scalar product (which is equivalent to the original one)
(x,Y)1 = L (xi,y;)1=1
where
X= Z; x,+x.,, y= Z; y,+ y.,, i=1
x,,y,EKer(U-X,I) for i= 1,2,...,p,
,_1
and
x,, Y., E./l. From this it follows that the operator U is stable. Theorem 3.9 enables us to make Theorem 2.8 more precise for the case of a J-unitary operator.
3 Invariant Semi-definite Subspaces
180
Theorem 3.12: Let U (EA+ U A_) be a J-unitary operator, let A be its non-unitary spectrum, A = A, U A2, A, fl A2 = 0, and let Az-' (X-' I X E A2) = Ai. Then the operator U has invariant subspaces J e± E such that the non-unitory spectra of U 12+ and U 19_ coincide with A, and A2 respectively. Moreover, A C p(U), and if X E A, (respectively, X E A2), then
the root subspace .If (UII+) (respectively, .x(UI If_)) coincides with -SPa(U). El
A C p(U) by virtue of Remark 2.4 and 2.Corollary 5.13. Let
If+=CLin(.)JU)IXEA1),
2'_=CLin(.If,,(U)IXEA2).
Since Al- ' f1Ai = 0 by hypothesis, .± C 1° by virtue of Exercise 7 on 2.§6. By
construction Y+ are completely invariant subspaces of the operator U and as(UIY+)=F1,(U)( x(UI2_)=Yx(U))when XEAl (respectively, XEA2). By Theorem 3.9 there are subspaces -ie7± E J(± which are invariant relative to
U, which contain .± respectively. Carrying out an argument similar to that used in proving Theorem _2.8 we realize that the non-unitary spectra of the operators U 12+ and U I _ consist of normal eigenvalues. By construction A, C a(U I I+) and A2 C a(U 12_ ), and the corresponding root lineals satisfy the requirements of the theorem. It remains only to notice that the skewconnectedness of .,,(U) and 2'x - ,(U) (see 2.Corollary 3.12) implies that
a(UI.+)nA2=0anda(UI2_)f1A,=0. In this paragraph we investigate the question of the number of invariant subspaces possessed by J-bi-non-contractive operators which have at least one 4
invariant maximal uniformly positive subspace (from not on the sets of uniformly definite subspaces from . C and ./l(- will be denoted by t!o and ..llo respectively). But first we prove the following Lemma 3.13: Let .e = M+ O+ operator of the subspace .ii'- E assertion holds: if the subspace
.
- be a J-space and let Q be the angular Then W = .,Y+ + /V- and the following
Y_(x++Kx+x+Ee+,K: Xe+
-W-, IIKII<°°)
and if 1 E p(QK), then
2=(y++Fy+Iy+E. +, F.
+- 'l-, II FII <1
and - 1 E p(QP- F), and
F= (P + Q)K(I+ - QK) Conversely, if
?=(y++Fy+Iy+E.,Y+,
IIFII
§3 Fixed points of linear fractional transformations
181
and - l E p (QP- F), then
Y= (x+ + Kx+ I x+ E .,Y+ , K: e+ -. ,W- ,
11 K11 < oo
and
1 E p(QK),
and
K = P- F(I+ + QP- F)-'. El
The equality
.
= . + + { "-
(3.4)
is obtained from 1.Corollary 8.16. Let
q'= (x++Kx+ x+E.W+, K:
IjKjj
Then, if x = x+ + Kx+ E Y, we have
x = x+ - QKx+ + Kx+ + QKx+ = (I+ - QK)x+ + (P- + Q)Kx+
= y+ + (P + Q)K(I+ - QK) y+ = y+ + Fy+, where
y + = (I+ - QK)x+
and F= (P + Q)K(I+ - QK)-'.
Therefore
Y _ ly+ + Fy+ I y+ E de+, F .
- .N-, 1 1 F 1 1 < oo J.
The assertion that -1 E p (QP- F) follows from the equality
QP F= QP(P +Q)K(I+-QK)-'=QK(I+-QK) =-I++(I+-QK)-i The other assertions of the lemma are proved similarly. Theorem 3.14: Let V be a J-bi-non-contractive operator, and let SP+ (E,41+ ) be an invariant subspace of V Then the operator V has an invariant subspace {.E -11-. If, moreover, V has an invariant subspace (9?+ # ).ii 1 E /tf+ such "that t'l + A= W, then this operator has not less than a continuum of
invariant subspaces in each of the sets ./tfo and /tf+\,-tfo .
JLet K+ be the angular operator of the subspace -V+. It follows from 1.Lemma 8.4 that II K+ II < 1, and by Lemma 2.2 Gv (K+) = 0. Therefore VE A+ and so, by Theorem 2.8, V has an invariant subspace A- E ..Jf-.
We now use results from 1.§7.6 and we shall suppose, without loss of generality, that 2+ _ ,Y+. Let Q be the angular operator of the subspace A"
and K, the angular operator of a subspace ./Vi (*.,Y+). By virtue of 1.Theorem 8.15 assertion c) 1 E p(QK,) and so it follows from Lemma 3.13 that.4"1+
= (y+ + Fly+J, where y+ E.W+ and F, = (P + Q)K,(I' - QK,)-'.
3 Invariant Semi-definite Subspaces
182
Moreover F, ;e 0. Since
+ and
V- are invariant relative to V, and
+ Fi: the subspaces . 'l G_' = (y+ + a F, y+ I y+ E W+) will also be invariant for all a E C. We note that the G+ are not necessarily semi-definite. {
But, for a sufficiently small in modulus we have - 1 E p(aQP- Ft) and therefore when, for example, I a I < z II F, II ', it follows by Lemma 3.13 that ,/I-Q+ =
(x++Kax+Ix+.
+, K.,=aP-F,(I++aQP-F,)-'} (3.5)
and since Ia1IIF1I1 <1 IIK.II< 1-a1IIF111
it follows that all such iV are uniformly positive. Moreover the power of the sets of them is not less than that of the continuum. Without loss of generality we can now suppose that J V E ,? , i.e., II K, II < 1 ;the equality IV; + V- = M is obtained from 1.Corollary 8.16). We consider the set of those a E C for which - 1 E p(aQP- F). This is an open set in each connected component of which the function a - K. is holomorphic. Let Z be the connected component of this set which contains the point of = 1 (see Lemma 3.13). We prove that there is a point ao E -Z such that II K.. II > 1. First of all we notice that if - P6 C, then Z has a finite boundary point CL'.
Then, if a E Z and a - a', we conclude from the form of the operator K. oo as n oo, and the existence of ao has been proved. But if ,?. = C and if, contrary to assumption; I K, II < 1 (a E C), then by Liouville's theorem K. = constant. Therefore QKq = aQP- F, (I+ + aQP- Ft)-' =
(see (3.5)) that II K,, II
I+ - (I+ + aQP- F1)-' = const., which implies the equality QP- F, = 0, and therefore K. = aP- F, 0 const. (because F, ;4 0)-we have obtained a contradiction. We now consider continuous functions
(a = )ff: [0, 1]
(0 < T < 1)
with
f,(0) = 1 and f,(1) = ao
for all rE [0, 1] and
A(t2) when T1 ;4 r2 and t1, t2E (0, 1).
Since all the functions fl(t) are continuous, the functions II Kf(t) II will also be
continuous with respect to t for each r, and moreover IIKf(o)II=IIK,II < 1 and
1,
and therefore there are t,E (0, 1) such that II KJ,(,,) II = 1. By hypothesis ;e-
when Ti ;4 T2. Therefore the operator V has not less than a
continuum of different invariant subspaces ./l "f (,_) E .,//+\. //o with the angular operator K,(,_).
§3 Fixed points of linear fractional transformations
183
In conclusion we shall show how to obtain from results about invariant subspaces for J-non-contractive operators similar results for operators acting in G-spaces. In 1§6.6 a construction was given for embedding a G-space in a J-space, in which, in particular, a G(')-space was embedded in IIX. We shall call such an embedding canonic, and we shall investigate what happens to certain properties of operators acting in the corresponding spaces. Later we shall use the following general theorem which we introduce here without proof. 5
Theorem 3.15 (M. Krein [2]): Suppose a Banach space R is densely and continuously embedded in a Hilbert space W. If a linear operator A: V - V with VA = 4 is continuous (respectively, completely continuous) in 4 and is also symmetric as an operator in Ye, then it is continuous (respectively, completely continuous) in ' and it can be extended by continuity on to the whole of .
into a self-adjoint operator A: ' e.
As a direct consequence of this theorem we obtain the following proposition which will be important later: Lemma 3.16:
Let the G-space.,Y = Ae® +- with G.± C W', and let .
be canonically embedded in a J space 7e _ + O+ - with ± canonically embedded in R±. If A with 9A = M is a G-selfadjoint continuous (respec-
tively, completely continuous) operator, then it extends uniquely into a continuous (respectively, completely continuous) J-selfadjoint operator A
R.
Let J (= I G I -G) be the unitary part of the polar decomposition of the operator G; in the given case J coincides with the difference of the ortho-
projectors P+ and P- on to+ and- respectively: J = P+ - P-. Then JA is a I G I -selfadjoint operator, and . - by construction (see 1.Proposition 6.14) In accordance with is the completion of i( relative to the norm (I G I x, x)1,12.
Theorem 3.15 the operator J extends into an operator J = P+ - P- (P`- are the ortho-projectors in .W on to ±), and JA extends into an operator JA selfadjoint in W which will be simultaneously with A continuous or completely continuous. Therefore A = Y JA is the required operator. Corollary 3.17: I f under the conditions o f Lemma 3.16 A = I I AU II' i=1 is a continuous G-selfadjoint operator and A 12, A21 E 9-, then A = II Au II ?i=1 where the A;i are the extensions of the operators A, (i, j = 1, 2) and moreover A12,A21 E .q',,, Since
A = Al + A2, where Al = II
All
A22 I
,
A2 =
0
A 12
A21
0
i
3 Invariant Semi-definite Subspaces
184
A, and A2 are G-selfadjoint operators, and A2 E Y., we have only to use Lemma 3.16.
Remark 3.18:
Suppose W + W+ O+ M- is a G-space, G.W± C e±, G± = G j .Y±, and 0 E p(G_ ). Then from A =Ac, 9A = .Ye and A12 E .y'0 it follows that A21 E 9'm. This follows from the equality A21 = G_'A12G+ which holds in this case.
If V`, together with V, is a continuous operator and then V and V` extend into continuous operators V and Vc in .,Y, and P = V`. Remark 3.19: v = 9 v, =
,
This follows immediately from Lemma 3.16 if we use the equality V= 2'(V+ V c) + i [ (V - V')12i] and the fact that the operators '(V+ V c) and 2
112i(V- V`) are G-selfadjoint. Theorem 3.20: Suppose . = W+ ( Ae- is a G-space, GW± C M±, G+ = G I .e±, and 0 E p(G_ ). If V is a G-bi-non-contractive operator (i.e., Vand
V` are continuous G-non-contractive operators, 9v = 9)V'= .'), V12 E 91., 9'+ C .?±, V9 -?+ = Y+, then there is an 9+ E ill+ (.e) containing q+ and such that V9+ C L+. Let M = . + O+ . - be canonically embedded in ie- = Se- + O+ .ie- -. It follows from 0 E p(G_) that.- = .e-. The operator V satisfies the conditions in Remark 3.19 and it therefore extends into a continuous operator V which is J-bi-non-contractive. Since V12 E 91m and 0 E p (G_ ), so V12 E Y.. Therefore, by Theorem 3.9, the operator V has the property -6 +. The condition V. + = Y+ implies the complete invariance of the closure of 2'+ in ' relative to the operator V. Consequently there is an ! = [ x+ = Kx+ I x+ E .3W+ } (E. #+ (.e)) containing
'+, and V2' =
It is easy to see that 2+ =.t fl '(E. lf+ (M)) will be the
required subspace.
Corollary 3.21: If Atis a Gt"1-space and if V is a G(')-bi-non-contractive operator, then any of its completely invariant non positive subspaces admits extension into a maximal non-positive invariant subspace.
We are in the conditions of Theorem 3.20 to within a change-over to the anti-space and taking account of the fact that V21 is a finite-dimensional operator. So V21EYm.
Exercises and problems I
Let V be a J-non-contractive operator. We define a generalized linear-fractional transformation F v' on a non-empty subset .7'(C.W + ): t V (.i) = Ul F V (K) I K E .J
.
§4 Invariant subspaces of a family of operators Prove that if
y3 = ( V) is
185
a semi-group of J-non-contractive operators, then
F'',j= (Fv I V E .Y31 is also a semi-group in relation to superposition of transformations and that F+vo o Fv2 = Fv, v,; if .13 is a group, then F. is a group, and 2
(Fv)-' = Fir Prove that if U is a J-unitary operator, then
11 K II < 1
Fu (K) < 1 and
11 Ku = I « I Fu(K) II = 1 (M. Krein and Shmul'yan [4], Potapov [1]). Hint: Use Proposition 3.1. 3
Let V be a J-non-contractive operator, U a J-unitary operator, and W = U-' VU. An operator Ko E k is a fixed point of the transformation F v+ if and only if Fij-(Ko) is a fixed point of the transformation F+w.
4
Prove that the function Fv generated by a J-non-contractive operator V E A_ has at least one fixed point in .W (I. Iokhvidov [10]). Hint: Use Proposition 3.2 and Theorem 3.9.
5
Suppose V is a J-bi-non-contractive operator, and F%' has a fixed point Ko with 11 Ko II < 1. Then the function Fv has either a single fixed point in .Yr+, or it has not less than a continuum of such points, and moreover not less than a continuum on the
boundary of the ball i+ (Azizov; cf. Khabkevich [5] ). Hint: Use the result of Problem 3 and suppose Ko = 0. Use Theorem 3.14. 6
Suppose TE A_ is a J-bi-non-contractive operator, (91+, '_) is a dual pair, Y'+ is completely invariant relative to T, and '_ is invariant relative to T`. Then there is an extension of this pair into a maximal dual pair with preservation of the properties mentioned (Azizov). Hint: Suppose T12 E .y'm and use 1. Corollary 8.24 and Theorem 3.8.
7
Prove that if U is a J-unitary operator, then the function Fu is continuous in the weak operator topology if and only if U12 E rm (Helton [3] ).
8
If VE A_ is a J-semi-unitary J-bi-non-contractive operator, then it has the property c (Azizov). Hint: Use Proposition 2.5 and Theorem 3.9.
9
Suppose V is a J-non-contractive operator, Y+ (C .?+) is its completely invariant
subspace with angular operator K+, and def Y+ < oo. Then V is a J-bi-noncontractive operator and there is a subspace Jv+ E ..tl+ which is completely invariant relative to it and which contains Y+ (Azizov). Hint: Verify that Fv is a closed transformation of the convex non-empty bicompact subset . V+ (Kr ), and use Theorem 3.6. 10
Suppose A E (L) is a J-selfadjoint operator, A is its non-real spectrum, A = A, U A2, A, = A2*, A, fl A2 = 0. Prove that if A is an A22-completely continuous operator, then the operator A has a maximal non-negative invariant subspace Y'+ C 9A such
that the non-real spectrum of the operator A 12+ coincides with A, (or with A2) (Langer [3], M. Krein [5]). Hint: Use Theorems 1.13 and 3.12.
§4 1
Invariant subspaces of a family of operators We have already encountered invariant subspaces of groups of J-unitary
operators in 2.0.3. From 2.Theorem 5.18 it follows, in particular, that if W = ( U) is a bounded amenable group of J-unitary operators, then without
3 Invariant Semi-definite Subspaces
186
loss of generality these operators can also be supposed to be unitary. So below,
instead of the condition of amenability and boundedness, we shall immediately impose on a group of J-unitary operators the condition that the operators entering into the group are unitary, i.e., U.W± _ .±. The following theorem makes 2.Theorem 5.18 more precise. Theorem 4.1: If W = ( U) is a group of operators which are simultaneously J-unitary and unitary, then it has the property 4) [1] .
Let ('+, &_) be a dual pair invariant relative to the group 40/, and let be its extension into a maximal invariant dual pair. Since the properties of being J-unitary and simultaneously unitary are -invariant (see Exercise 2 on § 1), we can suppose without loss of generality that (2+, 0-) is a
definite pair (see Exercise 6 on § 1). Since W'+ = k+, and all U E %?l are J-unitary, it follows that 6?l9+l1 = 2+1] . The duality of the pair (2'+, 2_ ) implies the inclusion 2'_ C 21+'1. Since (see 1.(10.1)) 2+11 = 9+ [(@ j ,I', (in the
present case the 4No which appears in 1.(10.1) is equal to [0)), and since all U E ?l are simultaneously J-unitary and unitary which implies Q/!2. + = 9 +, it follows that UN1 =,ii",. By virtue of 1.Theorem 10.2, in order to establish the inclusion 9- E It-, it is now sufficient to establish that TO_ is the non-positive subspace which is maximal in 2+1]. Let P, be the orthoprojector from .+11 on to .4i",. Since U22- = 9_ and U,,11'1 =,4 , we have UP,._ = P, L'_, U[(P,k_ )1 fl <4 i] = (P,9-) -L fl .ii",. The subspace and therefore . Consequently, (P,1- )1 fl i , (C JP+11) is negative and J-orthogonal to
T = Lin (9_, (P,'_ )1 fl +j) is the maximal negative subspace from 9+11 which is invariant relative to all UE 41. The maximality of_ as an invariant negative subspace from 91+'1 relative to the group Q?z implies the coincidence of _ with 21. That 2'+ E t?+ is verified similarly.
In this paragraph we investigate the question of invariant subspaces for groups of normal J-unitary operators. But first we prove some auxiliary 2
propositions.
Lemma 4.2: Let . ' be a J-space, U be a J-unitary and simultaneously positive operator, and let Ex be its spectral function. Then
a) 91, = C Lin (ExN' I X < 1)and2'2=CLin{(I-E,). Iµ> 1)areneutral mutually orthogonal subspaces; b) .0 = Ker(U - I) [ G) ] (2', G) Y2); c) J(SL', (1 2'2) = Y, O+
'2, J Ker(U - I) = Ker(U - I);
d) every definite invariant subspace of the operator U lies in Ker(U - I). a) Since Ea.
C
when X > 2', and 1 - E ).Yf C (I - E,,).*" when µ > µ,
§4 Invariant subspaces of a family of operators
187
to prove the neutrality of 2, and 2'2 it suffices to verify the neutrality of the subspace Ea.' when 21 < 1 and the subspace (I - E,,).' when µ > 1 respectively. From the spectral theory of self-adjoint operators it follows that or (U I E),.') C [ Xmin, X1, where Xmi,, = inf ((Ux, x) 11 x 11 = 1); moreover 0 E p (U)
because U is a J-unitary positive operator, and therefore Xmin > 0. The subspace E,,. is also invariant relative to the operator U` (= U-' ), and Since X<1 by a (U` I Ex. ) C [ 1 / X, I Xmin] hypothesis, so [Xmin, x] n [1/X, I Xmin] = 0. The neutrality of E,,.1f now follows from 2.Theorem 1.13. The neutrality of I = E,). when µ > 1 is verified similarly. The orthogonality of 2, and 22 follows from the orthogonality property of the
spectral function: Ex (I - E,) = 0 when X < and µ > 1. b) The orthogonality of Ker(U - I) to 2, and .?2 follows from the orthogonality of Ker(U - I) to EX.Y( when X < 1 and to (I - E,). when µ > 1.
We now verify that Ker(U - I) [1] 2, and Ker(U - I) [l] 2'2. Because of the continuity of the J-metric it suffices to verify that Ker(U - I) [1] Ea.' when X < 1 and that Ker(U - I) [1] (I - E,).W when µ > 1 respectively, but this in turn follows directly from 2.Theorem 1.13 when we take into account that a(U` I Ker(U - fl) = (1). We consider the subspace .' = Ker(U- I) [ S] (2, Q+ 22) which is invariant relative to U; we shall prove that W' = e. For otherwise we would have . _ .e' G) W ' 1 where ' 1 (7 0) is a subspace invariant relative to U, and U' = U I ,W" 1 is a positive operator. Let Ex' be its spectral function. It is easy to see that E,,.' 1 C E for all v E R. Therefore, since Ex.Y' 1 is orthogonal to E),.t (C 9?1 C Jr") when X < 1 and (I - E, ). Y' 1 is orthogonal to (I - E,, ).w' (C 22 C . ') when µ > 1, it follows that Ex ' 1 = (01 for X < 1
and (I- E,), ,Y' = (0) for µ > 1, and so a(V') _ (1). Hence we conclude that V' = I Ye ' 1 , which contradicts the orthogonality of W ' 1 to Ker(U - 1).
Consequently Ae' =. c) Taking into account 1.Formula (7.1), this follows from b). d) Let 2 be an invariant subspace of the operator U. It is well-known (see, e.g., [XXII] and cf. 4.Remark 1.8) that then ExS C 2. Since Eat is a neutral
subspace when X < 1, the definiteness of 2 implies Eat = (0). Similarly (I - E,)9? = (01 when µ > 1. Therefore X1(2, O+ 22), i.e., 2 C Ker(U - I). Corollary 4.3: Let the operator U satisfy the conditions of Lemma 4.2, and W defined on W let emu be the algebra of all continuous operators A: which commute with U. Then -Bu has a common non-trivial neutral invariant
subspace if and only if U;4 I. Each such subspace is J-orthogonal to Ker(U -
I).
If U = I, then Wu coincides with the algebra of all continuous operators
A: . -+ ., which, as is easy to see, has non-trivial (and including also neutral) invariant subspaces. But if U ;e I, then, for example the 2, and .?2 appearing in Lemma 4.2 are
188
3 Invariant Semi-definite Subspaces
neutral and invariant relative to -j6u, since each of the operators of this algebra commutes with Ex for all X E IR (see, e.g., [XXII] and cf 4Theoreml.5). Let 2 be any non-trivial neutral subspace invariant relative to .tiBu, and let P
be the J-orthogonal projector from ' on to Ker(U - I). Then, as is easily verified. P2' is an invariant subspace of the algebra u I Ker(U - I), which coincides with the algebra of all continuous operators acting in Ker(U- I).
Therefore either P2' = (0) or P2 = Ker(U - I). Since (P` = ) P E X u, so PY C 9 and therefore P-T is a neutral subspace. It follows from assertion b) in
Lemma 4.2 that Ker(U- I) is a projectionally complete subspace, and therefore PL = (0), i.e., 2'[±] Ker(U- I). The following lemma is of a general character and seems to be well-known.
Lemma 4.4: If Y = ( V) is a group consisting of normal operators and containing the conjugate V* whenever it contains V, then VU*U = U*UV for any U, V E Yl.
By hypothesis the operator W= V*VU*UE I' and therefore it is normal. Since W is a (U*U)-selfadjoint operator and the scalar product (U*U , ) is equivalent (in the sense of equivalence of the corresponding norms) to the original one, so a(W) C FR, which, taking the normality of W into account, is equivalent to its selfadjointness: W= W*, i.e., V*VU*U= U*UV*V. As is well-known (see. e.g., [XXII]) it follows from this that (V*V)12'2U*U= Let V= be the polar representation of the normal operator V; here S is a unitary operator commuting with U*U(V*V)1,12.
S(V*V)1,12
(V*V) 1/2. Since S(U*U)25-1 V*V= (V*V) 1/25(U*U)25- 1 (V* V) 112 = (VU*U)(U*UV)
= (U*UV*)(VU*U) = (U*U)(V*V)(U*U) = (U*U)2V*V,
so S(U*U)2 = (U*U)2S. We conclude, as above, that SU*U= U*US. Consequently VU*U= U*UV.
We turn now to the formulation and proof of the main result of this paragraph. Theorem 4.5: Let ,Y be a J-space and 4/ = { U) be a group consisting of
normal J-unitary operators and containing the operator U* whenever it contains U. Then =V! has the property
W.
Let (2'+, 2'-) be an invariant dual pair of the group ail, and let (2'+, ) be its extension into a maximal invariant dual pair. We use Exercise 6 on § 1 and we shall suppose that (2P+, 2'-) is a definite dual pair. By assertion d) of Lemma 4.2 we have 2'± C n{Ker(U*U- I) UE -V!). Moreover, the maximality of J+ implies the equality U*U = I for all U E W. For, if we had Uo*Uo ;d I for some Uo E -V/, then it would follow from Lemma 4.4 and
§4 Invariant subspaces of a family of operators
189
Corollary 4.3 that the group 4! has a common non-trivial neutral subspace
J-orthogonal to Ker(Uo Uo - I) and all the more to k+-so we have a contradiction. Thus U*U = I for all U E 4l, i.e., 4! is a group consisting of operators simultaneously J-unitary and unitary. By Theorem 4.1 it has the property 4) 1-1, and therefore 9± E -11:2:.
Corollary 4.6: If -V/ = ( U) is a commutative group consisting of normal J-unitary operators, then the minimal group containing 4l and W* has the property I']. The set 4!* = (U* I U E =R!) is also a commutative group consisting of
normal J-unitary operators. Moreover, if U, V E 4l, then UV= VU, and therefore by a well-known theorem of Fuglede U*V = VU*, i.e., the elements of the groups 4! and 41 commute with one another. Consequently, the group
FV generated by the union of 4l and 4l* is commutative and consists of normal J-unitary operators. Moreover, if UE 41, then U*E Ql. It only remains to use Theorem 4.5. The operator ball .yl+, as was shown in I. Proposition 8.20, is bicompact
3
in the weak operator topology, and therefore (see 1.Proposition 8.21) the centralized system of its closed subsets has a non-empty intersection-on this is based the proof of the following key proposition. Theorem 4.7:
Let Y' = ( V) be a family of J-bi-non-contractive operators, let F; = (Fv I V E Y ) be the corresponding linear fractional transformations of the ball JY (see (3.1)), and let .Ylb be the closed set of fixed points of the transformation F. If for each finite set (Vi) C 'F we have ni .X<, ;d 0, then n(.XirI VE Y' E Y') d 0. Corollary 4.8: Let Y = ( V) be a family of J-bi-non-contractive operators V with Vie E 9. (respectively, V21 E Y.). If each finite subset of Y has the then the whole family Y has the property D+ or 4 + (respectively, (D_ same property. If Y is a family of J-unitary operators with V1 2 E .gym (or, what is equivalent, V2i E
then similar assertions hold also regarding the properties 4'[11
and $110.
Let V be a J-bi-non-contractive operator, and let JYv be the set of fixed
points of the function F. By Theorem 3.8, when V12 E Y., this set J' is closed in the weak operator topology. By Proposition 3.4 each finite subset Y ' from Y has the property 4'+ if and only if n(.X v I V E Y -) # 0, and Y having
the property 4'+ is equivalent to n(.XvI VE Y') ;d 0. it only remains to use Theorem 4.7. The remaining assertions are proved similarly.
3 Invariant Semi-definite Subspaces
190
Corollary 4.9: If W= (U} is a commutative family of stable J-unitary operators and if U12 E 9. for all U E /!, then W has the property i [1] . Let U1, U2,. . ., U" be an arbitrary finite set of operators from i, and let Wn be the group generated by these operators. From the commutativity of W and the stability of the operators comprising W, we conclude (see II, items 5.17-5.20) that -Wn is a bounded amenable group. Therefore (see §4.1) -W has the property 4)[1]. It remains to apply Corollary 4.8. Under the conditions of Corollary 4.9 the whole family %1l is not necessarily bounded. The following example confirms this.
Example 4.10: Let I1 =1I+ O+ n- be a Pontryagin space, II+ = Lin(e+}, and II_ = C Lin{e-}i , (ei , e;) = Si; (i, j = 1, 2, ...), and let the sequence (;}i of complex numbers be such that Ei 1 tr 2 = 1, E; ;4 0 (i = 1, 2, ...). We consider the subspaces I
n
II In) = C Lin e+ +
Ete;, i=1
l
en-+2, ... }.
These too are Pontryagin spaces with x = 1, IIf") of"+'>, III")[-L] are negative subspaces, and II1 = of") [ + ] II(") [1] (n = 1, 2, ...). We define operators Un:
Unx = x when x E
of")'
Uny = - y when y E
It is easy to see that these are a-unitary stable operators (because U = I) and UnUn = U,nUn (n, m = 1, 2, ...). It can be seen from the construction of the operators Un that all their non-negative eigenvectors lie in III'), and therefore the non-negative eigenvectors of the family W = (Un) 1' lie in fl (IIf"t}; = Lin(e+ + E 1 ,e; ). Thefore JI/ has a single non-negative invariant subspace Lin (e+Ei E,e; } and it is neutral. But if the family ! were bounded, then the group generated by it would also be bounded. By 2.Theorem 5.18 this group, and therefore also the family Wl, would have to have a positive invariant subspace-a contradiction. 4 We pass on now to the investigation of the question of invariant subspaces for a-non-contractive operators. Theorem 4.11: Let Y = (V } be a commutative family of a-non-contractive operators in H,,. Then it has the property 4?.
In accordance with Corollary 4.8 it suffices to verify the assertion for a family Y consisting of a finite set of operators. We verify that such a Y has the property +. Let Y+ (C e°+) be an invariant subspace of this family, and let 9+ be its maximal invariant non-negative subspace containing M1+. We shall use Exercise 4 on § 1 and we shall suppose that 9+ is a positive subspace.
§4 Invariant subspaces of a family of operators
191
+ [+]+1', where 9+11 is again a Pontryagin space flx, with = x - dim Y+. We shall prove that x, = 0, i.e., that .'+ E ,u+.
Then II =
Suppose x, > 0 and let P be the a-orthogonal projector on to +1J . Then, as is easily verified, the operators V, = PV 12i+] form a commutative family y
, = ( V,), and moreover the V1 are 7r-non-contractive operators in
2P+11 (see
the more general Lemma 5.9 below. Since 9+ is the maximal invariant non-negative subspace of the family Y, it follows, on the one hand, that Y', has no non-negative invariant subspace 2, (B}: for otherwise Lin(9+,2,} would be an extension of 9+ into a non-negative invariant subspace of the family Y' and on the other hand, by Corollary 2.9, each of the operators V, has in Yl+'l a x,-dimensional non-negative invariant subspace. Moreover, if V, xo = Xxo, with (B ;4)xo E JO+, then I X I = 1. For, if I X I > 1, then, by Exercise 5 on 2.§6, 21,(V,) is a non-negative subspace and it is invariant relative to Y-which is impossible. But if we had I X I < 1, then (see Exercise 5 on 2.§6) xo E -, i.e., xo 410, and the isotropic part of 2a(V1) would again be invariant relative to y, which once more is impossible. Therefore, I X I = 1 and Ker( V, - XI) is non-degenerate. Consequently (see Exercise 23 on 2.§4)
the operator V, has no associated vectors corresponding to this X. Let X1, X2, ..., X, be all such points from a,(V,) to which correspond non-negative eigenvectors. It is clear that', = Lin (Ker( V, - X,I)) ° is a certain Pontryagin
space H. From Corollary 2.9 we conclude that x = xl. Moreover, ,)'1 is invariant relative to 'Vi. We consider the family 111 _ V1 01 = ( V I
, ). It
again is commutative and consists of 7r-non-contractive operators one
of which is, as is easily seen, a stable 7r-unitary operator. By carrying out the procedure indicated as many times as there are operators in Y/, we obtain a family 7i' = ( V') consisting of a-unitary operators which are the restrictions
of the original operators on to an invariant subspace IIx, with x' = x,. According to Corollary 4.9 the family '//' has a x1-dimensional non-negative subspace-we have obtained a contradiction. We now verify that the family Y has the property 4_ . Let Je_ be the maximal invariant non-positive subspace of the family, containing the original one. Again by virtue of the result of Exercise 4 on § 1 we can suppose that 1- is a negative subspace. Then 9-I'] is a Pontryagin space fl,', with x' = x and it is
invariant relative to the family Y" = (V` I V E Y } consisting of it-noncontractive operators. In accordance with what we have proved above, Y ` has a x-dimensional non-negative invariant subspace Y+ in *U1; but then ii'L+)
(E,11-) is invariant relative to Y and Y+> D 9-, and this is possible only when Yl+11 coincides with_ .
If ! = (U) is a family of pairwise commutating ir-unitary operators, then it has the property 4fll. Corollary 4.12:
Let (9+,.9'-) be the maximal invariant dual pair of the family W, containing the original one. In accordance with Exercise 6 on §1 we can
3 Invariant Semi-definite Subspaces
192
suppose that (2'+, 2'-) is a definite pair, and therefore [9+ [+19, ] [ll is a Pontryagin space rI,,, invariant relative to 1l with x' = x - dim 2'+ By Theorem 4.11 when x' > 0 the family 4! has in I1x' a x'-dimensional non-negative invariant subspace 2, and a maximal non-positive invariant subspace .'z= P;11 f H. Consequently (Lin(9+,.2,), Lin(Je1_,2'Z)) is an extension of the dual pair (2'+, 2-) preserving invariance relative to ill, and this is possible only when 91, = 2'2 = [0), i.e., k± E J1
.
Let .e be a G(")-space, let ail= (U) be a commutative family of G(')-unitary operators, and let 2'_ be its maximal invariant non-positive subspace. Then 2- E U. Corollary 4.13:
We use the results of §3.5 and canonically embed the G(')-space W in II.,
and we also extend the family 41 by continuity into a commutative family %1l = (U) of 7r-unitary operators. In accordance with Corollary 4.12 Wl has the
property c -, and therefore there is an 2 E - (ILL) which contains 2'_ and is invariant relative to TV. Then (cf. 1.Proposition 8.18) 2'fl E mil- O, 22 fl W' 3 2'-, and 2' fl W' is invariant relative to JIV. Since 2'_ is maximal it
follows that 2'_ = k fl Y. The result obtained enables us to prove a series of propositions about the existence of invariant subspaces; one of them is
Theorem 4.14: Let 4! = (U) be a commutative family of J-unitary operators, let (2+, 2-) bean invariant dual pair of 4l with def 2'+ < oo or def ,'_ < cc, and let (2'+, 2'-) be its extension into a maximal invariant dual pair. Then 2+ E It+.
Without loss of generality we shall suppose that def 2'+ < co, and so def 2_+ < co also. In accordance with Exercise 6 on § 1 we can suppose that
(2'+,2'-) is a definite dual pair. Consequently 9+11 is a G(')-space with x = def i+, the W 191+'1 are G('')-unitary operators, and 2' is their maximal
invariant non-positive subspace. We conclude from Corollary 4.13 that 2'_ E tl- (2'1+-LI), and so (cf. I.Theorem 10.2) 2'_E ll ( ). Therefore (x!11, 2'_) is an invariant dual pair containing (2'+, 9-), and since the latter is maximal we obtain 9+ = E ,tl+
Exercises and problems I
Investigate whether in Theorem 4.5 the condition (U E N!) _ (U* E -Y!) can be omitted.
2
Let Y = (V) be a commutative family consisting of a-non-contractive and ir-bi-nonexpansive operators. Then it has the property 4i (Azizov (6)).
3
Generalize Corollary 4.13 to the case where i+! _ (U) is a commutative family of G`-non-contractive operators and VU = VU = N' (Azizov).
§5 Operators of the classes H and K(H)
193
4
Prove that if -VI = ( Ul is a commutative family of J-bi-non-contractive operators, if 1'+ (respectively, 2'_) is a maximal completely invariant (respectively, maximal invariant) non-negative (respectively, non-positive) subspace of the family -f/ and def
5
Let W = Lin (e, fl, 11 e 11 = 11 f 11 = 1, (e, f) = 0. We introduce into . ' a J-metric by means of the operator J: J(ae + /3f) = /3e + af. Prove that the group generated by the operators J and U: U(ae+Of)= Xae+X-'(if (I a ;4 1) is soluble, consists of
'+ < co (respectively, def 2'_ < oo), then Y+ E fl+ (respectively, Y_ E Lf-) (Azizov).
jr-unitary operators, and has no common non-trivial invariant subspaces (cf. 2.Theorem 5.18) (Azizov). 6
Let d be a commutative algebra of operators acting in a J-space and closed relative to
the operations of conjugation and J-conjugation, i.e., A E d - A * E d, A` E d. Prove that if (Y+, °-) is any maximal dual pair invariant relative to d, then T± E ./ff± (Phillips [3] ). Hint: Use Theorems 1.13 and 4.5.
§5
Operators of the classes H and K(H)
In 1.§5.4 we introduced the concepts of the classes h± with which we shall operate in this section. 1
Definition 5.1: We shall say that a bounded operator T belongs to the class H
(TE H) if it has at least one pair of invariant subspaces 2+ E ..Zf+ and Y_ E -&- and every maximal semi-definite subspace Y± invariant relative to T belongs to MI respectively. From this definition and 2. Proposition 1.11 the implication
TEHa T`EH
(5.1)
follows immediately. We now investigate a number of other properties of operators of the class H.
Theorem 5.2: If an operator T has an invariant subspace of the class -it, fl h+ (respectively, .,L(- fl h-), then T E A+ (respectively, T E A_ ). In particular, T E H= T E A+ n A_ . D
Let K+ be the angular operator of the invariant subspace 99+ E
of the
operator T. If 2+ E h+, then K+ can be expressed in the form of a sum K+ = K1 + K2, where 11 K1 11 < 1, and Kz is a finite-dimensional partially isometric operator. By Lemma 2.2 Gi (K+) = 0, and therefore
GT(Ki )(K+-K2 )T>>+(K+-K2 )T12(K+-K2 )T21-T22(K+-K2 ) _ -K2+T11-K2 T12K+-K+T12K2+-K2+T12K2+ +T22Kz E.V and by Definition 2.3 TE A+. Similarly one proves that if the operator T has an invariant subspace _T- E , ll- fl h-, then TE A_. From Definition 5.1 and what has been proved, it follows that TE H = TEA+ n A-.
3 Invariant Semi-definite Subspaces
194
It follows from this theorem that the propositions proved earlier (see §§2-3) for operators of the class A+ hold also for operators of the class H. In particular, from Theorem 3.9 and Exercise 8 on §3 we obtain
Corollary 5.3:
If T (E H) is a J-bi-non-contractive operator, (respectively, a
J-semi-unitary J-bi-non-contractive operator or, in particular, a J-unitary operator), then T has the property c (respectively, Ill ). Corollary 5.4: Let T (EH) be a J-bi-non-contractive operator. Then each of its completely invariant non-negative (respectively, invariant non-positive) subspaces belonging to the class h+ (respectively, h-). It is sufficient to use Corollary 5.3, the Definition 5.1, and the simple fact that a subspace of a subspace of the class h± belongs to h+-.
Let T be a J-bi-non-contractive operator of the class H. Then
Corollary 5.5:
there is a constant xT < oo such that the dimension of each of the neutral invariant subspaces of the operator T does not exceed xT. Let (2') be the set of all neutral invariant subspaces of the operator T. It follows from Corollary 5.4 that dim 2 < oo for all ' E (i). If we assume that there is no constant XT < oo bounding the dimension of the subspaces Y, then with dim 9?;, oo, among them, there could be found a sequence dim 2;, <, dim We put Yi = Yi,
= Lin 191,,
(n = 2, 3, ...).
Y,;n
is a monotonely non-decreasing sequence of neutral invariant subspaces of the operator T, and dim 2 - oo as n - oo. But then is also a neutral invariant subspace of the operator T, and C dim C oo-we have obtained a contradiction. By construction
Corollary 5.6: Let T (EH) be a J-bi-non-contractive operator, and let Ker(T - XI) =,4'0\ [ + ],4'x+ [ + ] . l "X be the canonical decomposition of the kernel of the operator T- XI into an isotropic component A 'OX, a positive Then .4!"a are uniformly component .4-x+, and a negative component difinite subspaces and min{dim(.4'Q [+].-!"x; ),dim(.'l"0 [+].'!" )) <
From Corollary 5.4 we obtain that dim A"0 < oo, that ./l"a are uniformly definite subspaces, and that any definite subspaces in A "X [ + 1, /l "x are uniformly definite. Therefore by virtue of 1.Theorem 9.11 which .'! "a , dim .4) < oo, min (dim(.'!"0 [+].'!",; ), dim(.' U. [+]A "a )) < oo. min (dim
implies
the
inequality
Let T (XH) be a J-bi-non-contractive operator. If T has no neutral eigenvectors, then it has a single pair E #+ and 99- E -I of Lemma 5.7:
Y+
§5 Operators of the classes H and K(H)
195
invariant subspaces and they are uniformly definite, .' = 2+ -- 2-. Further-
more, if T is a J-semi-unitary or J-unitary operator, then 97 = Y+ [1] and
j'_9+[+]2'
.
Let 2± (E_#) be invariant subspaces of the operator T. By definition 5.1
T+ E h±, and therefore its isotropic part 9o is finite-dimensional and invariant relative to T. Therefore if T has no neutral eigenvectors, then Yo = (B], i.e., the subspaces Y+ are uniformly definite. By Theorem 3.14 and 1.Corollary 8.16 it follows that the 2'± are unique. Therefore if, in addition, T
is a J-semi-unitary operator, then TY+ = 2+ implies T`2' _ 9?+, and so TY+ [1] C 2+ I`]. Since the pair 2+ and 27 is unique we obtain 2- = Y+ [] It follows from 1.Corollary 8.6 that . _ 2+ + 27 or, if 2- _ 91+ [1], then
Lemma 5.8: For J-bi-non-contractive operators membership of the class H is a (D-invariant property.
Let V be a J-bi-non-contractive operator, let V E H, and let 2'o be a neutral invariant subspace of V. By Corollary 5.4 dim 20 < oo, and therefore by virtue of 1.Corollary4.14 V9?o = moo. It follows from the result of Exercise 20 on 2.§4 that the operator f ,induced by the operator Vin the factor-space . = X'119o is J-bi-non-contractive. It is necessary to prove that VE H. By Corollary 5.3 there are invariant subspaces 9± E ,e fl h ± of the operator V which contain 2'o. But then 22± _ 9+I9o E,/u±(.°) fl h± (cf. Exercises 21, 22 on 1. §8), and 2'± are invariant relative to V. Let V± E ,if1 ± (.°) be arbitrary invariant subspaces of the operator V. Then .N"± = (x± E R± I R+ E 4'+) are maximal semi-definite invariant subspaces of the operator V, and therefore .N± E h±, which implies that ;V+ (=.4'±/9o) are members of the classes h± respectively. By Definition 5.1 then, VE H. Let 2' be a uniformly definite invariant subspace of a J-bi-non-contractive operator V, and let V9 = 2' if 2 C ,+. Without loss of generality we suppose that 2' C+ or 22 C W- depending on the sign of Y. Then W _ Y [ O ] Y[1] and relative to this decomposition the operators V and J can be expressed in matrix form: V=II Vii II?i=1,
V21=0,
0Ep(Vii),
and
J= II JiJII?!=1,
J21=0,
12=O,
and
if
2'C.?
or
ll=-Il2if22C.?-;
J22=
J*_
l
3 Invariant Semi-definite Subspaces
196 and
22 = J* = J22'.
If 2E .10-, then
follows from Exercise 34 on 2.§4 that
it
V22
is a
J-bi-non-contractive operator. Suppose
2'C.+, X2E2'
,
and
=-V-'V12x2.
x1
Then [ V22x2, V22x21 = [ V(xi + X2), V(xi + X2)] i [XI + X2, xi + X21
= [xl, xi] + [X2, x2] i [x2, x21,
i.e., V22 is a J22-non-contractive operator. Since 2[l] is invariant relative to
the J-non-contractive operator V` it follows that (V22)` (=(V`)22) is a J22-non-contractive operator, i.e., V22 is a J22-bi-non-contractive operator.
Thus, in both cases-whether 9 C ,JP + or 9 C 9--the operator V22 is J-bi-non-contractive.
Now suppose, moreover, that V E H. We verify that then V22 E H also. Suppose, for example, that 9 C :+, and that 2+ (E./tf+ ), existing by virtue of Corollary 5.3, is an invariant subspace of the operator V containing 9, and that 9+ = 9 [+] 9. .9 is completely invariant relative to V, so V2292 = 22 . It follows from 2+ E /ff + fl h+ that 9 E +ww) fl h+, i.e., the operator V22 has at least one invariant subspace from -&+ (91,W) fl h+. We now assume that V22,4-2 = T2+, and ./U= Lin[2'l.Az J. Then
A z E.alt+and
( . W ) and V,11'= IV. By Definition 5.1 A' E h+ and so -Y? E h+ also, i.e., all maximal non-negative invariant subspaces of the operator V22 belong to the class h+.
By Theorem 5.2 V22 E A+ and therefore V22 has at least one maximal non-positive invariant subspace (in 91±] and therefore also in .Y). Let 2- be such a subspace. Then 9+ = Y- 41 n Ytl1 (E.11Z+211) is an invariant subspace of the operator V`. Hence 2+ E h+, which implies (see Exercise 12 on 1.§8) that 9- E h-, i.e., V22 E H. A similar assertion in the case 2' C ?- is proved in the same way. Thus we have proved Lemma 5.9: Let 2' be a uniformly definite invariant subspace of a J-bi-noncontractive operator V, with V2 = 2' if 2' E let P, be the J-orthoprojector from . on to 9, and let V22 = (I - P') V(I - Pr) (I - Pr)'. Then V22 is an 40+,
(I - P,)*J(I - P,)-bi-non-contractive operator, and if V E H, then V22 E H also.
Definition 5.10: We shall say that a family of operators .1'= [ X] which commute with a J-bi-non-contractive operator Vo of the class H (i.e., 2
§5 Operators of the classes H and K(H)
197
XVo J VoX), with p(X) fl C+ it 0 for every X E ?' belongs to the class K(H) and we shall write E K(H). We remark that ;t' may consist of a single operator X and we shall then say that the operator X belongs to the class K(H) (X E K(K )). As well as the notation 2" E K(H) we shall also use the symbols -E K(H, Vo) to show precisely with which operator of the class H it is that .' commutes. Lemma 5.11: Let I,= (V ] be a commutative family of J-bi-non-contractive operators with 1' E K(H, Vo). Then when J * I (respectively, J ;d - I) the family 'o = 7/ U Vo has a non-trivial completely invariant non-negative subspace _T+ (respectively, invariant non-positive subspace Y_).
We consider first the case when the operator Vo has no neutral eigenvectors. Then, in accordance with Lemma 5.7, there is a single pair Y+ E .W and Y- E A 1 of subspaces invariant relative to Vo, and VoY+ = Y+ It
follows from the uniqueness of Y+ and 2- that Vf+ = Y+ and V2- C 2for any VE 11, i.e., 11/o2+ = 2+, 'Yo Y- C -T-. We put T+ = Y± We now assume that the operator Vo has at least one neutral eigenvector xo: Voxo = Xoxo. Then Ker(Vo - X01) is an invariant subspace of the family V. If
Ker(Vo - XoI) is degenerate, then, in accordance with Corollary 5.6, its isotropic part 4'00 is finite-dimensional. We consider the three possible cases:
a) IX01 < 1; b) IXoI> 1; c) IXo1. a) I X0 < 1. In accordance with Exercise 6 on 2.§6, Ker( Vo - XoI) is a
non-positive subspace, and therefore (see Exercise 17 on 2.§4) .4't is invariant relative to each of the operators VE Y Moreover, by "Q0 2.Proposition 4.14, V.,4"Oo = A "0,, since dim , l < oo. We put Y+ _ 9- = .410. b) I Xo I > 1. In accordance with Exercise 5 on 2.§6, Ker(Vo - XoI) is a non-negative subspace. It follows from Exercise 17 on 2.§4 that .%1.O is the
isotropic part of Ker [ Vo - (l/Xo)I] and therefore V`.iV°x. = 1 0o for all V E T. We again use Exercise 17 on 2.§4 and we obtain that V41 = ol '°o for all V E Y' We again put Y+ = Y_ = ,11'L c) 1 ao 1 = 1. with 19 2.§4, In accordance Exercise on Ker( Vo - XoI) = Ker( Vo - )%ol), and therefore Ker( Vo - X0I) is invariant relative to both V and V` for all VE 1'. Consequently, if yo E At, then [ Vyo, x] = [yo, Vo x] = 0 for all x E Ker( Vo - X01), i.e., V.4'00 is a subspace
in VO,. From 2.Proposition 4.14 and the fact that
.4"0o is
finite-
dimensional, we obtain that V.NO0 = 4'°, for all V E Y . Again we put We now assume that Ker(Vo - X01) is non-degenerate. In accordance with Corollary 5.6 Ker( Vo - XoI) is a Pontryagin space. From the consideration of the cases a)-c) we see that I Xo I = 1, and the family Y induces in Ker (Vo - XoI)
198
3 Invariant Semi-definite Subspaces
bi-non-contractive operators relative to the form [ , ] I Ker( Vo - )bI). By virtue of Exercise 2 on §4 we obtain that these operators have common non-trivial semi-definite invariant subspaces 2'+ and 2'_, and that (VI Ker(Vo-XoI))J'+=2'+. Theorem 5.12: Let a commutative family Y = ( V) of J-bi-non-contractive operators belong to the class K(H, Vo). Then the family Y o = Y U Vo has the property 4). Since the property (a family of J-bi-non-contractive operators is commutative) is 4?-invariant, it follows from Lemma 5.8 that the property (Y o belongs to the class K(H )) is also 4?-invariant. It therefore suffices to prove that every maximal completely invariant subspace 2'+ C _P++ U (B) and every maximal
invariant subspace 2'_ C .40 -- U (0) of the family i'o belong to tl+ and -lfrespectively. Let 2' be one of such subspaces. By Corollary 5.4 2' is uniformly definite, and therefore N' = 2' [--] 211" . let Pr- be the orthoprojector from . on to Y. We consider the family Y 02 = ( Vzz) U (V0)22 of V- Pr-)*J(I - Pr)-binon-contractive operators of the class K(H, (Vo)ZZ) (see Lemma 5.9), where
V22 = (I - P,)V(I - PA I
Yf11,
(Vo)22 = (I- P,')Vo(I - P1) I
x,[11
From the maximality of 2' and Lemma 5.1 applied to the family X02 we conclude that 9?[11 has the sign opposite to that of Y. In accordance with 1.Proposition 1.25 2 is a maximal semi-definite subspace. Corollary 5.13: A J-bi-non-contractive operator V of the class K(H) belongs to A+ fl A_ and has the property.$. This follows immediately from Theorems 5.12 and 5.2.
3
Let a J-bi-non-contractive operator V belong to the class K(H, Vo), Yo be the maximal invariant neutral subspace of the operators Vo and V, and let 2'+(E,//+) be invariant subspaces of these operators containing 2'o. Let 2+ = 2'o [ + y' _ 2' [ O ] 99_ be the decompositions of the subspaces 2+ into the sum of a neutral subspace Yo and uniformly definite subspaces 2±. Since 2+ E //+ fl h± we have ?o11 = Lin (2+, 2- ), and therefore, by virtue of the formula 1. (7.4) and 1.Corollary 8.16 "Y = 2'0 0+ (2'+ 4-
'-) O J?o.
(5.2)
Relative to this decomposition the operator V is represented by the matrix
V=
Vii
V12
V13
V14
0
V22
0
V24
0
0 0
V33
V34
0
V44
0
(5.3)
§5 Operators of the classes H and K(H)
199
We mention that in a number of cases later (for example, in the proof of Proposition 5.14) we shall make use of the condition VE K(H ) in order to have the expansion (5.3), and therefore it would be possible to demand the existence of the expansion (5.3) instead of the condition VE K(H ). However, in order to avoid awkwardness of presentation we shall not do this. 5.14 Let J-bi-non-contractive operator V belong to K(H), and let e E ap( V) fl T. Then the root lineal 99A V) is a subspace, and it splits up into a
J-orthogonal sum 2e( V) = Nf [ + ] .Jfe, where dim ./V, < oo, VA', = .ib"ei . Ife C Ker(V - cI) and -ffe is a projectionally complete subspace or, in particular, lfe = (0). We consider the decomposition (5.3) of the operator V. The subspace lfe = tee( V) fl Ker V12 + f(V) fl Ker V13 is regular and moreover i = 1, 2, 3. The operator V22 is non-contractive Vie( V) fl Ker V,; C Y4
relative to the scalar product [x, y] (x, y E 2+ ), and V33 is non-expansive relative to the scalar product - [x, y] (x, y E ?_ ). Therefore (cf. Exercise 23 Ker( V;, - cIi), i = 2, 3, which is equivalent in the present on 2.§4) _T,( case to the inclusion .ite C Ker(V- eI). Since by virtue of Exercise 19 on 2.§4 /f, C Ker(Vc - eI), and qp(V) = V, [+] .ife, where .Ne =X21] fl Y,( V), it follows that .-/V, is invariant relative to the operator V. By construction dire [1] Ye(V) fl Ker Vl; (i = 2, 3).
It remains to verify that IV, is finite-dimensional, from which it will also follow that it is completely invariant. We assume the contrary: '4', is an infinite-dimensional subspace, and ?o, 2'+ and Y_ are the same as in the decomposition (5.2). Since dim(.?o O+ J-Wo) < ao, so a', fl ('+ + Y_) is an infinite-dimensional subspace, and therefore at least one of the subspaces fl Y_ will be infinite-dimensional. Suppose, for example, .'1 fl _r+ or that dim Fe fl Y+ = oo Since dim 2+Ker V,2 < oo, it follows that .4l"f fl Ker V12 ;e (0), and we obtain a contradiction with the fact that .4l", [1] 2e(V) fl Ker V12.
From Corollary 5.6 and Proposition 5.14 we obtain Corollary 5.15: It under the conditions of Proposition 5.14 the operator V belongs to H, then, when dim Y4 V) = co, .,lff, is a Pontryagin space with a finite number of either negative or positive squares.
4 We now consider a family 1= (A) of maximal J-dissipative operators of the class K(H) and we investigate for it the questions of the existence of maximal semi-definite invariant subspaces. Theorem 5.16: Let .4 = (A) be a set of maximal J-dissipative operators, let
.-c/EK(H, Vo), and let .'+ (C.?+) be invariant subspaces of the family ,V U Vo with Y+ completely invariant relative to Vo; and suppose there is a
3 Invariant Semi-definite Subspaces
200
point XA E p(A) fl C+ such that XA, XA E p(A 1 9?+) and XA E p(A I 9_ ). Then,
if the resolvents of the operators A E ' commute pairwise, there are extensions ?+ E .,tl± of the subspaces 2'± which have the same properties.
D We consider the family of operators r = (V = Kx(A) I A E d). In accordance with 2.Theorem 6.13 the V are J-bi-non-contractive operators. It follows from the inclusion 4 E K(H, Vo) that r' E K(H, Vo). Since XA, XA E p (A I '+ ), the subspace 2'+ is completely invariant relative to 'V, and the condition XA E p(A I T_) implies the invariance of 2'_ relative to 1. Consequently, by virtue of _Theorem 5.12, there are k+ E ill± fl h- containing U Vo. We now Y+ and such that VV+ = 2'+, V2'- C 2'_ for all V E I o = prove that k+ are invariant relative to .d, and that XA, XA E p(A I 2'+) and
XA E p (A I 2'_ ). To do this we note that 2'o = .'+ n 2- is an invariant finite-dimensional neutral subspace of the family 11, and therefore, by virtue of Exercise 13 on 2.§6, Yo c InVA I A Ed) and .d2'o C 2'o. We bring into consideration the i-space = 2'011/2'o and the families of operators acting in it gel = (A) and Y'o = (V) U Vo generated by the families and 1/0. Moreover, lA = (qA n 201")/2'o, XA E p(A) and KXA(A) = KxA(A). We note that because k± are maximal semi-definite the equality (.+ + 2'_ ) [1] = 2'o follows, and since 2± E h±, so (see Exercise 19 on 1.§8) 2011 = 2'+ + ,'_, and therefore
/2'o + 2'_ /1o. Relative to this decomposition of f the operators FE 'To have the diagonal form
Y=
V,1
0
0
V22
and therefore
- ( AK-' A
which implies that -ql+
)-
V
II
KXA'(VI I) 0
0
KXA'(V22)II'
are invariant relative told and that XA and XA belong
to p(A I 2'+/2o) and XA belongs to p(A 2'_/2o). It only remains to use the fact that XA, XA E p(A 12'o).
This theorem enables us to prove the following assertion which will be useful later.
Theorem 5.17:
Let A E K(H, Vo) be a maximal J-dissipative operator, let
be a J-space, where Kero(A - XI) is the isotropic part of the subspace Ker(A - XI), and let Vo be
X = X E pp(A ),
let .W' = Ker(A - X)/Kero(A - XI)
the J-bi-non-contractive operator induced in .' by the operator Vo. Then VoEH.
In. accordance with 2.Corollary 2.17 Ker(A - XI) = Ker(A ` - XI), and therefore Ker(A - XI) is a common invariant subspace of the operators Vo
and V. Hence, it follows, in particular, that Kero(A - ceI) also is their common completely invariant subspace. We conclude from 2.Theorem 1.17 that Vo is a f bi-non-contractive operator. We now verify that Vo E H. Since
§5 Operators of the classes H and K(H)
201
Vo E H it follows that it is sufficient to find at least one maximal non-negative
invariant subspace of the operator Vo. Let Y+ be the common maximal non-negative invariant subspace of the operators A and Vo (we know that there is such a subspace from Theorem 5.16). We consider the subspace P), = Y+ fl Ker(A - XI). We first assume, and later prove, that the deficiency (see 2.Theorem 4.15) def -Cvx of the subspace
'x = Lin(2\ Kero(A - XI)J/Kero(A - XI) is finite. By virtue of Exercise 4 on §4 the operator Vo has at least one maximal non-negative invariant subspace, and therefore, as proved above, Vo E H. It remains to verify that def _4 < oo. We consider the decomposition of the root subspace 2' (A) = 4'x [+] / , where /t"x and /lh, are invariant relative to A, dim ,/t'x < co, and M), C Ker(A - XI). The existence of such a decomposi-
tion follows from Proposition 5.14 and 2.Theorem 6.13. We assume that def 2a = co. Since dim ,/VX < co, there is an xo E Ker(A - XI) which
J-orthogonal to 2a(A) fl Y+. In accordance with 2.Corollary 2.18 xo is also J-orthogonal to all the subspaces 2',(A) fl Y. Let Yo be the isotropic part of the subspace 2+. Since dim Yo < oo, we have
is
We consider the 2 0 C Lin[X,(A) fl Y+ j it E a(A)], and therefore J,-subspace , = Lin[Y+, xo)lYo . Since 2+ E h + and xo 0 Y+, it follows that .°i is Pontryagin space with one negative square. Since 20 and Lin[,?+, xoI 9,1.'
are invariant relative to A it follows that A induces in, a J1-dissipative operator A. Since 20 is the linear envelope of the isotropic parts of the subspaces ? (A) fl 2'+(µE ap(A)), the root subspaces of the operator A are
positive-which contradicts Corollary 2.9 and Theorem 1.13, according to which the operator A must have at least one non-positive non-zero eigenvector.
Corollary 5.18:
Under the conditions of Theorem 5.17, if Ker(A - XI) is not semi-definite, then the operator A I Ker(A - XI) has semi-definite invariant subspaces Ya of both signs, which are maximal in Ker(A - XI). Moreover,
for all X = except possibly a finite number of points we have !a + 4- = Ker(A - XI), and for the remaining X E P we have dimKer(A - XI)/2' ); + 91Z < co.
Let Ker(A - XI) be not semi-definite, let !a be the semi-definite invariant subspaces, maximal in e, of an operator Vo belonging to the class H, and let 2a be their generating subspaces in Ker(A - XI). The maximality of .-a in .)Y implies the maximality of YK in Ker(A - XI), and the invariance of ._a
relative to Vo implies the invariance of 9a relative to Vo. Sincea E h±, it follows (see Exercise 19 on I.§8) that dim dim M fl a < oo, which in turn implies that (dim .i(/2 + jla =) dim Ker(A - XI)/ft + 91a < oo.
From the J-orthogonality of Ker(A - X1) and Ker(A - µl) when X µ, X, µ E ER, and from their invariance relative to Vo, Vo (EH ), it follows that only
3 Invariant Semi-definite Subspaces
202
a finite number of the subspaces Ker(A - XI) and sa can be degenerate. Moreover, again from the fact that Y,; E h- we obtain that dim Ker(A - XI)/f)+, + 91a = dim Ya n z- - dim Ko(A - XI), and the equalities
dim Ker(A - XI)/2 'a + £a = 0 and -Ta + Ya = Ker(A - XI) are equivalent.
5
Let j11= [ U) be a commutative group of J-unitary operators, and
1 E K(H, Vo). If Vo is also a J-unitary operator, then, by carrying out an argument similar to that used in the proof of Theorem 5.12, we can satisfy ourselves that the family -lo = %' U Vo has the property i) ['l. But if Vo is not a J-unitary operator, we can at present assert only that the family -11o has the property . However, in any case the group W does have the property JlJ . For, let Y+=moo[+]Y+E..rr+n h+ (WO C.-Po,c+C ++U[O)) be an invariant subspace of the group X71. Then Y_
=,+ [iJ
= So [+].S_ E _rr- n ..*,-
is also an invariant subspace of this group. Without loss of generality we can suppose that 2'+ C W± and £'o J. (2+ [ @] S-) Therefore in the decomposition (5.2) of the sRace ye all the components are pairwise orthogonal to one another and (moo + JIo) [1] (Ib+ [ (@] Y_ ). Let ./V+ (C P ±) be arbitrary maximal invariant subspaces of the group %1, with A + [1] /I"_, i.e., A'-) is a dual pair. We verify that .4'+ E To do this we consider the subspaces .4'± = ! "± n 2o1J. This too is a dual pair, invariant relative to X11, 1"± and dim I /I"± < dim A < co. Together with will (,/I"+' , also be an invariant dual pair if /I"+' = Lin [ 'o, .4 + ). We consider the f-space _ 0 IYo and the group '11 induced in it by W. Since Ib± C ,Y±, we have = I+ [ O] 2-, 11 + = J'±, and therefore %1l is a group of operators which are simultaneously J-unitary and unitary. By Theorem 4.1 '11 has the property J1J, and therefore (see Exercise 7 on 1.§5) there is an ..!"+ E,11+ such that is a dual pair invariant relative to 1, and moreover /I"+' C,, j7+. The subspace A", = Lin [.A"+, .'I "+) is invariant relative to -V1, is J-orthogonal to A" and contains at least one maximal non-negative subspace (/l "+ ). By construction it is a W(')-space with x (
quently (cf. 1.Theorem 10.2 and I.Proposition 8.18) every non-negative subspace which is maximal in A', is also maximal non-negative in -Y. _ We verify that there is in ./l ", a maximal non-negative subspace containing .4"+ and invariant relative to W. To do this it suffices to note that in ./I.+
the G(' -space .4{ ', = ,l -,/.'! "o (where .A! "o is the isotropic part of ./1', ) the group
X11,, induced in 4", by the group -11 1 A",, has (by Corollary 4.13) a maximal
§5 Operators of the classes H and K(H)
203
non-negative invariant subspace containing Lin(N+, /V_ ), and then to make use of the second part of the proof of 1.Theorem 10.2. and dim .A"_/.iV_ < co it follows that Since .il"+ C .A [l "[1] < co. Since (IF, n 4"I1.A _) is an invariant dual pair of
dim,il'+n
the group
! and contains
def /V+ < oo. By Theorem 4.14 Thus we have proved
A'_), we have A"+ n 4 11 = .4 + and A"±
E U-.
Theorem 5.19: A commutative group of J-unitary operators belonging to the class K (H) has the property4 . In this paragraph we describe the structure of the spectra of J-unitary and J-self-adjoint operators of the class K(H ).
6
Let U be a J-unitary operator and U E K (H, Vo ), and let x," = max(dim `moo I Yo C .00, Vo2o = Yo}. Then the non-unitary spectrum of
Theorem 5.20:
the operator U consists of a finite number x, (>,0) of pairs of normal eigenvalues disposed as mirror-images relative to the unit circle T, and ap(U) n T contains a finite number x2 (,>0) of points to which correspond the non-prime elementary divisors, and x, + X2 < xv
By virtue of Corollary 5.13 U E A + n A_ , and from Remark 2.4 and 2.Corollary 5.13 we conclude that the non-unitary spectrum of the operator U consists of normal eigenvalues. We use Exercise 6 on 2.§6 and obtain that 21, = Lin(Ya(U) I I X I > } is a neutral lineal. Since Vo2', C 91,, we have dim 2'i = xi < xv0. Since the spectrum of a J-unitary operator is symmetric relative to the unit circle, the first part of the theorem has been proved. Let -9 = (e E T I UI Y, (U) has non-prime elementary divisors). Then, by virtue of Exercise 14 on 2. §6, Ker(U - el) is a degenerate subspace; let Y0 denote its isotropic part. Since Ker(U - el) is invariant relative to Vo and Vo, we have VoYof! = Yo, and therefore '2 ° Lin (Y° I C E e) is a neutral lineal invariant relative to Vo. Consequently dim Y2 = x If < x v,,, and so f consists of a finite number x2 of points and x2 < x2'. Since Lin(2,, Y2) is also a neutral subspace invariant relative to Vo, we have x, + x2 < x i + xz < X V,,Corollary 5.21: x, ,
Let A be a J-self-adjoint operator and A E K(H, Vo), and let
be the same as in Theorem 5.20. Then the non-real spectrum of the
operator A consists of a finite number x, (,> 0) pairs of normal eigen values symmetrically situated relative to IR, and ap(A) n IR contains a finite number x2 (>, 0) of points to which correspond to non-prime elementary divisors, and x, + x2 < xV,,-
It follows from the definition of the class K(H) that the Cayley-Neyman
transform K,,(A) of an operator A belongs, when X E p(A), to the class
3 Invariant Semi-definite Subspaces
204
K(H, Vo) whenever A does. It remains only to apply Theorem 5.20 and II.Corollary 6.18. Thus, if a,, X1; X2, X2; . . . Xx,, Xx, are the different non-real eigenvalues of a J-self-adjoint operator A E K(H ), then "I
.f = [+] (Ya; (A) + _1rs,(A )) [+] 4e',
(5.4)
i=i
where Y' is the J-orthogonal complement of E;,k , [+] (14 (A) + 1a;(A)), and the operator A is expressed relative to (5.4) by a diagonal matrix which we write symbolically as x,
A = [+](A),;+Aa) [+]A',
(5.5)
i=1
where A,, = A I Y, (A), and A' = A I .W', and moreover the spectrum of the operator A' is real, and a(A,,) _ [µ) (µ = X1, X,; X2, X2; ...; Similar expansions ,W=
[+](I%,(U)+tee,i(U))[+]
(5.6)
i=1 x,
U = [+] (Ue; + Ue, ') [+] U'
(5.7)
also hold for J-unitary operators U of the class K(H ), where [ei, e;
the different non-unitary eigenvalues of the operator
are all
U,
E [ ei, ti-'), and the spectrum a(U') is unitary.
In particular, the expansion (5.6) and (5.7) are valid also for J-unitary operators of the class H. 7 In conclusion K (H ).
we give some examples of operators of the classes H and
Example 5.22: Every J-non-contractive or J-bi-non-expansive operator in Ux belongs to the class H (cf. Definition 5.1 with Corollary 2.9).
V be a J-bi-non-contractive operator and let V = V' + V", where V" = I I V; I I;;=, is a diagonal operator relative to the .,Y-; let p (vii) fl P0/12) = 0, and V" E Y.. Then decomposition Example 5.23: Let
VE H.
We conclude from the representation V = V' + V" that V E A+ fl A_ . E7 Consequently, by virtue of Theorem 2.8, it has at least one pair &+ and !- of maximal semi-definite invariant subspaces. Let K± be the angular operators of these subspaces. In accordance with Lemma 2.2
K+ V - V22K+ = - G v°(K+) and
V;1K- - K- V22 = G V-- (K- )
§5 Operators of the classes H and K (H)
205
Since p(vii) n P (V22) = 0, we obtain from [VI] that these equations can be rewritten in the equivalent form
K+= -
(V2'2 - XI-)-'G v.'(K+)(Vii - µI+)+'
1
47r
K-
(Vl' -µI+) 'G = - 47r12
r,
µ-
r,..
dX dµ
(5.8)
)(Vz2 - XI-)-' dX dµ
(5.9)
where IF v;, and r v j, are non-intersecting Jordon contours surrounding a( VI l ) and a(V'2) respectively. Since V" E .99., so G:.. (K±) E 9 9., and therefore we
obtain from the equations (5.8) and (5.9) that K± E 91m. It only remains to
notice (see Exercise 18 on I.§8) that this is sufficient for the inclusions 97+ E h±.
Example 5.24: An amenable uniformly bounded group W of J-unitary operators belongs to the class K (H ). As the operator Vo with which all the operators U E Qi will commute we may take (see Exercise 1 below) Vo = p+ - p-, where P+ are the Jorthoprojectors on to the components of the maximal uniformly definite dual
pair (97+, 97-) invariant relative to °?l (the existence of such a pair is guaranteed by II.Theorem 5.18).
Exercises and problems 1
Verify that the operator JE H.
2
Prove that every uniformly J-bi-expansive operator belongs to the class H. Hint: Use Theorem 2.1.
3
Prove that a J-unitary operator U belongs to the class H if and only if there is a canonical decomposition W' = ,Yl [+] .,Y1- such that the angular operators (with respect to this decomposition) of all invariant subspaces 99 ± E //± of the operator U are completely continuous (Azizov [9]).
4
Prove that, under the conditions of Proposition 5.14, if V E K (H) and if di (f), d2(f), ... , dr,(e) are the orders of the elementary divisors of the operator VI -We, then [di (f ) eET i=t
where x,.,,
2
is the same as in Corollary 5.5. In particular, if V is a J-unitary
operator, then r,
di(f)
eET i=i
(Azizov: cf. Usuyatsova [2] ).
2
x>1
3 Invariant Semi-definite Subspaces
206
Hint: Similarly to that was done in proving 2.Theorem 2.26 show that there is in M' (V) a E;=, ([ d;(c)]/2)-dimensional neutral subspace invariant relative to V0. In the case of a J-unitary operator use Theorem 5.20. 5
Formulate and prove an analogue of Proposition 5.14 and Exercise 4 for maximal J-dissipative operators and, in particular, for J-self-adjoint operators of the class K(H) (for the latter taking Corollary 5.21 into account).
6
Prove that is a J-bi-non-contractive operator Vo belongs to H, and if Yo is its finite-dimensional
non-degenerate
invariant
subspace,
then
the
operator
Vo I 91J1J E H (Azizov).
Hint: Use Lemmas 5.8 and 5.9. 7
Prove that in the expansion (5.7) U' E K(H) (Azizov). Hint: Verify that if UE K(H, Vo), then Vo.Yt" C .W', and use the result of Exercise 6.
8
Prove that if Uo E H is a J-unitary operator and if all its root subspaces 2'((U) with eI=1
are non-degenerate, then the space ."' in the expansion (5.6) can be
represented as the J-orthogonal sum of subspaces invariant relative to Uo and any operator community with it: k
.W' _ [+] 4;(Uo)[+]
(5.10)
"
where the ('r are the eigenvalues of the operator Uo to which corresponds at least
one non-trivial neutral vector (i = 1, 2, ... , k), and W" is the J-orthogonal complement to the first k terms of the expansion (5.10) (Azizov). Hint: Use Theorem 5.20 and Corollary 5.15. 9
Let eW = ( A) E K (H) be a family of maximal J-dissipative operators, let a/ E (L)
and the resolvents of the operators of this family commute pairwise, and let 1/'+ c p+ n (n VA I A E . /) be an invariant subspace of the family ._d. Prove that then there is a subspace 5C'+ D Y+ : 9+ E ..It+, 9+ C (n 1'A I A E,V), d9+ C
'+
(Azizov [9] ).
Hint: Use Theorems 1.13 and 5.12. 10
Let :-V = (B) be a commutative family of J-bi-non-contractive operators containing an operator Bo E H, and let W = (A) be a family of J-bi-non-contractive operators which has the property that AB = BA for any A E ', BE -V and for arbitrary A 1, A2 E . / there is a B12E . such that Al A2 = B12AZA I. Prove that then the family ..el U .-3 has the property 44 (Azizov and Dragileva [1]).
11
Prove that a metabelian group of J-unitary operators of the class K(H) has the property 4i [11 (Azizov; see [IV]).
12
Let Y = (V) be a scalarly commutative family (i.e., for every V1, V2 E Y there is a constant X12 such that V1 V2 = >112 V2 V1) of J-bi-non-contractive operators containing
an operator Vo E H. Then Y has the property 4 (Azizov). Hint: Use the scheme of the proof of Theorem 5.12. 13
Prove that if, under the conditions of Exercise 12,
Y
is a group of J-unitary
operators, then it has the property 4 L ] (Azizov [9], Khatskevich [4], cf. Helton [2] ). 14
Prove that a bounded J-selfadjoint operator A belongs to the class K(H) if and only if it has at least one invariant subspace 9+ E ,/!+ n h+ (Azizov [13] ).
Remarks and bibliographical indications on Chapter 3 15
207
Let A E K(H) be a bounded J-selfadjoint operator, and let 2' be a finite-dimensional non-degenerate subspace invariant relative to A. Prove that A I 2' E K (H) (Azizov).
Hint: Let M'+ E .:#+ fl h+, AY+ C 2+. Verify that det(2+ fl 2'W) < oo, use Theorem 4.14 and Exercise 14. 16
Prove that if V is a J-bi-non-contractive operator of the class H, if Se is
its
projectionally complete invariant subspace, and if the operator V 191 has a maximal (in 2') completely invariant non-negative subspace 2+, then V 199 E H (Azizov). Hint: Use Corollary 5.4 and Theorems 5.2 and 2.8. 17
Let the conditions of Problem 16 hold, and let P be the J-orthoprojector on to Y111 . Prove that the operator PV I M'1' E H and that it is P*JP-bi-non-contractive (Azizov).
Hint: Consider the subspace 2'
E tl+ : 2'+ C 2+ V2+ = 2+ and prove that P1'+ E h+, (PV 12111 )P2+ = P2+ and that V` I 2'tll is a P * JP-bi-non-contractive operator.
Remarks and bibliographical indication on chapter 3 §1.1-1.3. In the formulation of Problems 1.2, 1.3, 1.6, 1.7, 1.10, and 1.11 their sources were indicated. To what was said there it should be added that S. L. Sobolev proved that a 7r-self-adjoint operator in II1 has a non-negative eigenvector. The examples 1.4 and 1.8 were constructed by Azizov. J-selfadjoint operators A with J' C iA were first considered by Langer [2], [3] (see Definition 1.5). The Remark 1.12 is also due to him (see Langer [13]. The reduction of Problem 1.10 to Problem 1.11 was borrowed from Phillips [3]. §1.4. Theorem 1.13 was proved by Azizov [9] §1.5-1.6. The definitions and propositions are due to Azizov. We mention that the idea of formulating Lemma 1.20 in terms of `a complex J (' (see also 4.Lemma2.5 below) was borroed from [XI]. §2.1. Theorem 2.1 is due to Azizov (see Azizov and Khoroshavin [1]). The
existence of the property c for a uniformly J-bi-expansive operator was proved earlier by a different method in [VI]. Similar results for doubly strict focusing plus-operators (cf. 2.§4) in the cases of real and complex J-spaces by the methods of compressed mappings in the metric mentioned in the Exericses on 1.§8 were obtained by Khatskevitch [6], [11], [15], A. V. Sobolev and Khatskevich [1], [2], and in the case of real II earlier by Krasnosel'skiy and A. V. Sobolev [1]. §2.2. Lemma 2.2 was established by M. Krein [4]. Regarding Definition 2.3, see Azizov and Khoroshavin [1] and also Azizov and Kondras [1]. In the last paper the Remark 2.4 was also made, a remark stimulated by Noel's article [1].
§2.3. Proposition 2.6 was borrowed from Langer (see [XVI] ). Theorem 2.8 is due to Azizov and Khoroshavich [1]. Its first part was essentially proved earlier by I. Iokhvidov [10]. The remaining assertions are encountered in particular cases and in parts earlier in the papers of M. Krein [4], [5], [XIV], Brodskiy [1], Langer [l]-[3], Azizov [4]-[8], Azizov and E. Iokhvidov [1],
208
3 Invariant Semi-definite Subspaces
Azizov and Usvyatsova [1], I. Iokhvidov, M. Krein, and Langer [XVI]. The
same also applies to Corollary 2.9. We remark that our proof of these propositions is based on the application of a device of M. Krein [4] of approximating to a J-non-contractive operator V by uniformly J-expansive operators VI,,. Corollary 2.10 was formulated by Azizov. §3.1. Krein-Shmul'yan linear-fractional transformations were first brought into consideration by M. Krein [4]. They and their generalisation were systematically studied in the articles of M. Krein and Shmul'yan [4], Shmul'yan [2], [6], [8], Larionov [4], Helton [1]-[3], Khatskevich [5], [13], [14], and others. Propositions 3.1 and 3.4 are due to M. Krein [4], [5]. §3.2. Theorem 3.6 plays in our exposition the same role as was played earlier in these questions by Tikhonov's principle and its generalization given by I. Iokhvidov [10]. Theorem 3.8 is a natural generalization of the corresponding propositions of M. Krein [5] and I. Iokhvidov [10]. Its simple proof given here, based on Glicksberg's thgeorem, is by Azizov. §3.3. Theorem 3.9-a direct consequence of Theorem 3.8-was formulated the first part was proved by Wittstock [1] and, essentially, earlier by I. Iokhvidov [10]; the second and third parts were obtained by M. Krein [5] (cf. Langer [3] ). Corollary 3.10-see M. Krein [5], I. Iokhvidov [10], Whittstock [1], and Brodskiy [1]. Corollary 3.11 was noted by Azizov. Theorem 3.12 is due to Langer [3], M. Krein [5]; the short proof of it in the text is due to Azizov. §3.4. Lemma 3.13 and Theorem 3.14 were proved by Azizov. Earlier results close to Theorem 3.14 were published by Larionov [4] and Khatskevich [5]. §3.5. Theorem 3.20 and Corollary 3.21 follow from Wittstock's results [1]; Azizov gave the proofs given in the text. §4.1. Theorem 4.1 is due to Phillips [3]; the proof of it given in the text is due to Azizov. §4.2. The results in this section are due to Azizov. Lemma 4.4 seems to be well-known. Theorem 4.5 in the case where W is commutative constitutes the main results of Phillip's work [3]. As Shul'man noted, Corollary 4.6 is also true without the requirement for %' to be commutative. §4.3. The idea of applying Theorem 4.7 in the questions examined here is due to Naymark [2]. Esentially Corollary 4.8 is also due to him. Corollary 4.9 was obtained by Larionov [2], [3]. Example 4.10 was constructed by Azizov. §4.4, §4.5. Theorem 4.11 was proved by Azizov [6]. Corollary 4.12 was proved earlier by Naymark [2]. Corollary 4.13 is due to Azizov. Theorem 4.14 is also due to him. and proved here by Azizov. In the case when V12 E .9
§5.1-5.4. Helton in his article [2] considered a commutative group of J-unitary operators containing an operator Uo which has at least one maximal
non-negative invariant subspace and the angular operator of each such subspace is completely continuous. This provided the stimulus for the introduction by Azizov [9] of operators of the class H and K(H ). All the results of these sections, except Proposition 5.14, are due to Azizov. Propo-
Remarks and bibliographical indications on Chapter 3
209
sition 5.14 was mentioned in Usvyatsova's paper [2]; the proof in the text was given by Azizov. §5.5. Theorem 5.19 (see also Exercise 13 on this section), which is due to Azizov [9] and Khatskevich [4], generalizes a corresponding result of Helton [2]; in this paper the property 4 AI is proved by a different method. §5.6-5.7. The results in these sections are due to Azizov.
4 SPECTRAL TOPICS AND SOME
APPLICATIONS
In § 1 the concept of a J-spectral function is introduced and some classes of operators for which it exists are distinguished. Some of its applications to the solution of problems of invariant subspaces for families of `definitizable'
operators and to the construction of the (J1, J2)-polar decomposition of (Jl, J2)-bi-non-expansive operators are also indicated. §2 deals with questions of completeness and basicity of systems of eigenvectors and principal vectors of J-dissipative and, in particular, J-selfadjoint operators. In §3 we give some examples and applications.
§1
The spectral function Let E be a homomorphism mapping a ring e?, generated by certain
1
intervals of the real axis and containing the interval (- oo, oo ), into a set of
J-selfadjoint projectors E(A) acting in a J-space ., and let E(0) = 0; E(-oo,oo)=I; E(Af10')=E(0)E/0'); E(DUA')=E(0)+E(0') if 0 fl 0' = 0 (0, 0' E W). We denote the carrier of this homomorphism by a(E). A point X E a(E) is called a point of definite (positive or negative) type if
there is an interval 0 E 41 containing X and such that E(A).Y( C ,?+ or E(A).Y( C .40 - respectively. The sets of points of positive and negative type are
denoted by a+ (E) and a- (E) respectively. Let s = {ai )1 be a finite set of real points, and let 9t = .91(s) be the ring generated by the intervals for which points from s are not boundary points. Definition 1.1:
The homomorphism E described above, defined on 5?(s) is
called a J-spectral function with a set s = s(E) of critical points if, when 210
§1 The spectral function
211
Xo, µo 0 s(E), µo < Xo, and µ 1 µoi the strong limit of the operators E((µ, X0] ). exists and it coincides with E((µo, Xo] ). Definition
1.2: A point a E s(E) is called a regular critical point of
a J-spectral function E if the strong limits s-limx ,,E((- oo, X]) and s-limaiaE((X, oo)) exist. Otherwise a is called a singular critical point. If s(E)
contains no singular critical points, then E is called a regular J-spectral function.
Definition 1.3: A J-spectral function E with a set s(E) of critical points is called an eigen spectral function of a J-selfadjoint operator A if:
1) AE(A) = E(0)A; 2) a(A I E(A)M) C 0; 3) when 0 E 91(s) and o fl s(E) = 0 we have AE(0) = Jnt dE,, where the integral converges in norm if A is finite, and strongly if 0 is infinite. The set s(E) is called the set of critical points of the operator A and is denoted
by s(A), i.e., s(E) = s(A). 2 We now indicate some classes of J-selfadjoint operators which have a J-spectral function. One of them is the class of operators K(H) and, in particular, H (see Exercise I below). `Definitizable operators' form another class.
Definition 1.4: A J-selfadjoint operator A with p(A) ;d 0 is said to be definitizable if there is a polynomial p(X) such that [p(A)x, x] > 0 when X E cAk, where k is the degree of the polynomial. In particular we observe that a J-non-negative operator A with p (A) 0 is definitized by the polynomial p(X) = X. We shall not stop now to prove that definitizable operators have a J-spectral function; this result is formulated later as an Exercise. Instead we shall prove the key theorem in the theory of definitizable operators.
Theorem 1.5 Let A be a bounded J-non-negative operator. Then it is possible to set up in correspondence with every real number X ;d 0 one and only one J-selfadjoint projector EE, such that the function X E,, satisfies the following conditions: 1) if X 5 it, then ExE,, = EEx = Ea;
2) if X <,u < 0, then [Eax, x] > [Ex, x], and if µ > X > 0, then [E,,x, x] <, [Ex, x] for all x E *9;
3) ifx< -11A1l, then E,,=0, and ifX Ail, then E),=T; 4) if X ;d 0, then the strong limit s-lim,,lxE = E),+o exists and coincides with Ex;
5) If T is a bounded operator which commutes with A, then TEa = Ex T-,
4 Spectral Topics and Some Applications
212
a(A I E>,.) spectrum 6) the a(A I(I- EE). ) C [X, co).
lies
(-oo, X]
in
and
v dE2, converges in the strong operator topology as an Moreover, JIIAII -IIAII o with a singular point X = 0, and the operator improper integral S=_ A - JIIIIAIII - o
v dEE is a bounded J-non-negative operator, and S2 = 0,
SE), = E,,S = 0 when X < 0, and S(T - E),) = (T - E>,)S = 0 when X > 0.
We put B = JA. This is a non-negative operator, and therefore the operator C = B 1/2 JB 1/2 = C* satisfying the condition CB 1/2 = B 1/2A is 1/2 112
properly defined. Since II C II 5 II B sion of the selfadjoint operator
= II B II = II A II, the spectral expanbe written in the form
C can
is the spectral function, continuous C = 11-I IIIIAII - ox d F,,, where (F,) on the right, of the operator C. We put C,, = C I Fi,M when X < 0, and Ca = C I (I - Fx).f when X > 0, and E>, = JB1/2C 1 FF,B'/2 when X < 0 and I - Ea= JB1/2Ca'(I- F),)B1/2 when X> 0. 1) The operators Ex, are bounded for every X # 0. Moreover, if X < 0 Fx (respectively, Ca' (I - F)')) are selfadjoint
(respectively, X > 0), then
operators, and therefore the Ex are J-selfadjoint
with X ;4 0. Since
CCx ' Fx = Fa when X < 0, and CCa ' (I - Fx) = I - F>, when X > 0, it follows that ExE,, = EE), = Ex when X <,u. Hence, in particular, we conclude that the Ex are projectors.
2) If X
1
FB 1/2x, B 1/2x) - (Ca 1 F,B 1/2x, B 1/zx) f µ 1 dFa(B'/2x, B112x) -Jav
0
for every xE e.
Similarly it can be verified that [EE,x, x] < [Ex, x] when 0 < X < µ. Assertion 3) follows directly from the definition of the function X - Ea. 4) Let X <,u < 0. Then II E,,x- Eax ll2 = II JB112(Cµ' F,, - Ca 1 Fi,)B112x II2
II B II :
d(F,,B1/zx, B1/zx) 5 II B II v2
-
II(Fµ - Fa)B1/2x 11 2,
and since by hypothesis (Fxj--- is a function continuous on the right, s = lim1,1E,, = E. The assertion is verified similarly when 0 < X < µ.
5) Let T be a continuous operator and let TA = AT. We verify that TEa = E1,T when X < 0. By definition, E1, = JB'/2 j''- 1A11 -0 1/v dFFB1/2 when
X < 0. Let 2 (Pn)' be a sequence of polynomials with
sup (Pn(X)IXE [-IIAII,IIAII)<M
and
lim pn(v)=1 if -IIAII
n-w
n-ao
§1 The spectral function
213
Then
s- lira JB'/2pn(C)BV'12 = E. n-m Since
JB112C"'B112 = A"'+' (m=0,1,2 ....), we have TJB''2pn(C)B1/2
= JB1/2pn(C)B1/2T, and therefore TES, = E1, T. Similarly one verifies that TEa = E),T when X > 0. In particular, AEI, = ExA.
6) We observe that, for any bounded operators T1 and T2 the sets a(TiT2) U (0) and a(T2T1) U (0) coincide. This follows immediately from the formula (T2T1 - XI)
(T2(T1T2 -
XI)-'T1 - I)
which holds when 0 * X E p (Ti T2 ). Consequently, since
AE1,= JB'i2F',Bv2 and
A(I-E),)=
JB1,12(I-
F)JB1,12
(X ;4 0)
we have
a(AEx) U (0) = a(B1/2JB1/2F),) U (0) = a(CFx) U (0), and
a(A(I - Ex)) U (0) = a(B1/2JB''2(I - Fi,)) U (0) = a(C(I - F,)) U (0). Hence
a(A I Ea.') C (0) U (- oo, X], a(A I (I- E),), W) C (0) U [X, co).
It remains to verify that if X < 0, then 0 0 a(A I E), ,Y), and if X > 0, then 00 a(A I(I - Ea).W). We shall verify the first of these statements; the second is
verified similarly. Suppose that X < 0 and 0 E a(A Ex'). Since A I Eat is a non-negative operator relative to the form [ , ] I E), and E),.*' is a Krein space, it follows by virtue of 2.Corollary 3.26 that a,(A I E),,W) = 0, and therefore 0 E o,, (A I E),.-W) U a, (A I E>,.W). In both cases there is a sequence (xn) E such that 11 Eaxn 11 = 1 But then and AE>,xn - 0. BV2AEExn = (Bv2E',X = C,FXB'12Xn - 0 also. Consequently Exxn 0 also-we have a contradiction.
We now verify that the function X -+ E), satisfying the conditions 1)-6) is uniquely defined. Let X - Ex' be another such function with X < µ < 0 or 0< X < µ. It is clear that E a = E;E,+ E),(I - En). By condition 5) the operators E, and E,, commute. Consequently, Ex(I - E,) is a J-orthogonal projector on to the subspace Ei.f fl (I - E,,).W which is invariant relative to A. By condition 6) we obtain a(A I Ea. fl (I - Eµ) ) C (- co, x] fl [µ, oo ) =0, i.e., Eat fl (I - E,,).Jf = (B), and therefore Ea = E,;E,,. Similarly E,, = ExE,. When µ 1 X we conclude that Ei = E,\E>, = E',E) = E) i.e., we have
proved 1)-6).
4 Spectral Topics and Some Applications
214
Now let c > 0. Since e
e
J -IIAII-0
v dE,,= JB'/2C
'
J-IIAII-0
v dFFBtn = JB"2C_,F- eBvz = JB1/2
PBiiz,
and IIAII
IIAII
v dE, _ -
f
e
v d(I - E,,) = - JB'/2Cf '
v d(I - FF)B'/z
e
e
= JB'nCC
IAII
I
f
I
IAll
P FB' 2 = JB1/2 CC 'CfI - FF)Bvz e
= JB'12(I- FF)Bin = A - JB'nFeBvz, it follows that 11
11
P dE, = A - JB'1z(Fo - F_o)B'/z -IIAII-0 We put S = JB'iz(Fo - F_o)B1/2. Since Fo - F_o is a projector on to Ker C, we have
Sz = JBvz(Fo - F-o)Bvz JB1 z(F0 - F-o)B1/2 = JBvz(F0 - F-o)C(Fo - F_o)Bin = 0. Similarly,
SE),=ExS=0 when X<0 and S(I-F>,)
=(I-E),)S=0 when X>0. Corollary 1.6: A bounded J-non-negative operator A has a J-spectral function E with a single critical point X = 0. Moreover,
a(E)\(0) = a+(E) U a_ (E) and a+(E) = (0, oo) fl a(E), a- (E) = (- oo, 0) fl a(E).
As E(A), when 0 = (a, l], a 0, 0 96 0, we put E(i) = Ea - Ea. The fact that this determines a J-spectral function of the operator A with a single critical point X = 0 can be seen from the properties of the function Ex investigated in Theorem 1.5. In particular, it follows from properties 2) and a+(E) = (0, oo) fl a(E), a(E)\(0) = a+ (E) U a_(E), and 3) that
a- (E) = (- oo, 0) fl a(E). Remark 1.7: It follows from condition 5) of Theorem 1.5 that if T is a bounded operator which commutes with A, then it also commutes with S.
Remark 1.8: Since (see the proof of property 5) in Theorem 1.5) the
§1 The spectral function
215
operators E, (X ;d 0) are the strong limits of certain polynomials of A, every subspace 2 which is invariant relative to A is invariant relative to Ea and S. Moreover, ' = E,,/ [ + ] (I - EE,)!, E 2' and (I - E,)2' are invariant relative to A, and a(A I E),?) C (- oo, X], and a(A I (1- Ea).T] C [X, co). If, moreover, ? is invariant relative to a bounded operator T which commutes with A, then E),2', (I - EE,)Y and S9? are also invariant relative to this operator.
We note also that by virtue of condition 3) in Theorem 1.5 the integral m p dE,. In future, where it is f II - o p dE. can be written in the form I% necessary to do so, the operators E, and S corresponding to the operator A II
will be denoted by the symbols EEA) and SA.
Corollary 1.9:
If A and D are commutating J-non-negative operators, then
E> A)Ea°) = Ea°)EX(A); SAES;°) = Ea°DSA = O when X < 0 and SA(I (I_ E)(°)SA = O when X > 0; SASD = SDSA = O; DSA = SAD = 0.
The permutability of the operators ExA), EaD), SA and SD follows from El condition 5) in Theorem 1.5 and Remark 1.7. In accordance with Exercise 2 on
2.§2, WsA is a neutral lineal. On the other hand, we conclude from the commutativity of the operators SA and EaD) when X 71 0 that it is invariant relative to EX(D). Consequently Ea°)JsA is a negative lineal when X < 0, and (I- EX(°")?sA is a positive lineal when X > 0 (see Theorem 1.5 condition 2)).
Hence it follows that E!°I is, _ (B) when X < 0 and (I_ E)(°)) 3 s, _ (0) when X > 0, i.e., Ex°DSA = 0 when X < 0 and (I - Ea°))SA = 0 when X > 0. By virtue of condition S in Theorem 1.5 the operators D and SA commute, and therefore the lineal ?sA is invariant relative to D. Since Y"sA C -,ip °, it follows from the inequality I [DSAx, Y]12 < [DSAx, SAx] [Dy, y] = 0 that DSA=0. It remains to verify the equality SASD = 0. Since
VE,(°) = s-lim (DEYD + D(I - Ef' )), we have flo
We now use the fact DSA = 0 and SD = D - f
p dEv°)SA = 0.
p
Let ./ = (A) be a commutative family of bounded J-selfadjoint operators. We shall say that it has the property 'J if every invariant dual pair of this family admits extension into a maximal dual pair invariant relative to .,d.
3
Later in Exercise 8
it
will be proved that a family of definitizable
J-selfadjoint operators has the property (i(11 if this family is either finite, or if for all the operator entering into it the `corners' P+ AP- I W- are completely continuous. But here now we shall prove the validity of the following result. Theorem 1.10: If d = (A) is a commutative family of bounded J-nonnegative operators, then d has the property 4) [l1 .
4 Spectral Topics and Some Applications
216
Let (2'+, 9-) be an invariant dual pair of the family .1. We have to prove +
that there are Y+ E Jf such that Y+ D Y+, Y+ [1] 2-, and l2' C Y+. We first verify this assertion for the case when the Y+ are definite. Let
f+(A)=C Lin((I-EaA))l a>0}, f_(A)=C Lin{EaAt.I X <0}, Lin(f+(A)IAE.1), and 'V=C Lin(91s.,IAEd) It follows from Corollary 1.9 that f + (,d) [1] 6'_ (. /) and /V [1] 41 + (.d), and from
Remark 1.8 that '+ C (f + (,t))[1[. By virtue of this same Remark 1.8, SAY± C Y+ (A E V). But since the '+ are definite and MA, C
°, we have
SAY+ = (B} (A E .z), and therefore q+ [1] .u{
From what has
We consider the subspaces ?'+ = C Lin (Y+, f +
been proved above it follows that 2+ [1] 2_. We verify the inclusions Y+' C e + . In accordance with Exercise 11 on II. §3 the subspaces +
that .4 "[1] C Lin(99+,f+(,RI)}. Thus (Y', Y-' ) is an invariant dual pair of the family d and it contains (2'+, Y_). We Let (2+, 2-) be an arbitrary maximal dual pair containing + By construction for all A E , '. shall prove that Al C Y± )+ C Y, 1 [1 C 6 _ (A) [1[ and therefore AY+ C A6'-(A)1'1. Since C
Lin (Y+, f +.j7)) C JP
A = SA + f
.
It
remains
to
use
the
fact
v dE(PA), we have
Ae_(A)[1[ C C Lin(J1s,,,6+(A)}C Y+ C Y+. Consequently, AY+ C Y+. Similarly AY- C 2- (A E d). Thus the theorem is proved when 2'+ are definite. Let (9+, Y_ ) be an arbitrary dual pair invariant relative to .,l, and let VY)=C Lin(Y+,Y+ n.9°}. Then (Y(+'), YIP) is an extension of the pair (Y'+, Y_), and Y1+) [1] 241), Ac+) n A(l) = Y4) n Y(+)[ll = Y,() n i W[L] _ Yo,
and dYT C YQI). We pass to the consideration of the factor-space e= Yd1[14, and of the family of Jnon-negative operators It = (A} induced in it by the family e1, where I' 1) = YTIY0. In and its invariant dual definite pair accordance with what has been proved above, the family . ,/ has an invariant
maximal dual pair (+, 7-) containing (.+), j'T). In accordance with Exercise 7 on L§5 the subspaces 7 + = (x + E f c± I X± E dual pair in .,Y, and C Y+ D Y+.
i +) form a maximal
4 In conclusion we show how the results in §1.2 can be applied to obtain a (JI, J2)-polar decomposition of a (J1, J2)-bi-non-expansive operator. But first we give a definition.
Definition 1.11: A bounded J1-selfadjoint operator R is called a J1-module of a bounded operator T acting from a J1-space .01 into a J2-space . 2 if c(R) C [0, oo), R2 = T'T and Ker R = Ker T`T. Theorem 1.12: Every (J1, I2)-bi-non-expansive operator V has at least one J1-module R and admits a `(J1, J2) -polar decomposition' V = WR, where W
is a (J1, J2)-bi-non-expansive operator which is simultaneously a (J1, J2)semi-unitary operator or is conjugate to a (J1, J2)-semi-unitary operator.
§1 The spectral junction
217
E The operator A = I, - V` V is a bounded Ji-non-negative operator. Since from 2.Theorem 4.30 and 2.Remark 4.29 we have a(V` V) C [0, co), it follows
that a(A) C
1). Let Do = (Z, 1], and let E(go) be the corresponding value of the J1-spectral function of the operator A. From the definition of the J1-spectral function it follows that E(Do), 1 and (I, - E(Do)), 1 are invariant
relative to the operator V` V, and moreover a(V`VI E(Do).1) C [02] and a(V`V (I, - E(Do)).Y,) C [Z, 11 V`V 11]. In accordance with Corollary 1.6 the subspace E(Do)1 is uniformly positive. Since V` VI E(Do)W1 is an operator
which is selfadjoint relative to the form [ , ], I E(Do)1 and has a non-negative spectrum, it is non-negative and therefore it has a non-negative square root R1. Let F be a Jordan contour, enclosing the interval [Z, 11 V` V111, symmetric relative to IR and not containing X = 0 within itself, and let ,X be an
analytic function taking positive values for positive X. Then (see [XXII]) the
operator R2 = -(I/2iri)f r;X(V`VI (I, - E(Do)). 1 - XI1) ' dX is bounded,
= V`Vj(I, - E(Do)),J'1, a(R2) C [1/,2,jj V`V11], and, because IF is symmetric relative to R, the operator R2 will be J1-Hermitian. Since R2
E(Do).W1 [ + ] (I, - E(Do))Y1 = MI, the operator R such that Rx = Rix if x E E(Do)J'1 and Rx = R2x if x E (Il - E(Ao))J"1 is J1-selfajoint and bounded, R2 = V` V, Ker R = Ker V` V, a(R) C [0, oo), i.e., R is the J1module of the operator V. Vi = V I E(Do)4e1 and We now consider the operators is V2 = V I (I, - E(Do)) 1. Since E(Do)de, uniformly positive, (I, - ElDo)).1 contains the uniformly negative subspace which is maximal in
.i'1, and therefore V(I1 - E(Ao))., contains the uniformly negative subspace maximal in W2. From 1. Corollary 8.14 we conclude that .?Iv, (= v,) is projectionally complete, and MW is uniformly positive. Since E(Do).k1 and (I, - E(Do)).1 are invariant relative to V`V it follows that V` V,l C E(Do) 1, V` v, C (Ii - E(Do)), 1, and ,v, C XV,1. The operator V, as an operator acting from the Hilbert space E(Do)W1 with the scalar product [ , ] 1 into the Hilbert space 91IZl with the scalar product ]2 has the decomposition V1 = W1R1, where W1:E(Ao),,Y1-,W1`1 is [ semi-unitary or is adjoint to a semi-unitary operator. It is now sufficient to put Wx= W,x if XE E(Do)Mi, and Wx= V2R2- 'x if xE (Il - E(Do)). 1, and we obtain that V = WR and W satisfies the requirements of the theorem.
Exercises and problems 1
2
Prove that J-selfadjoint operators A of the class K(H) with a real spectrum have a unique J-spectral function with a finite number of critical points I X, , where X, are the eigenvalues of the operator A with a degenerate Ker(A - ail) (Azizov). Hint: Use the result of Exercise 15 on 3.§5 and the decompositon analogous to 3.(5.3) for a bounded J-selfadjoint operator of the class K(H). Similarly to Definition 1.3 introduce the concept of an eigen J-spectral function of a
J-unitary operator and prove that J-unitary operators UE K(H) have a
J-spectral function, which is, moreover, unique (Azizov).
4 Spectral Topics and Some Applications
218 3
Let Il, be a space with x negative (respectively, positive) squares, let 9' be a x-dimensional non-positive (respectively, non-negative) invariant subspace of a a-selfadjoint operator A, and let k,, X2.... , X, be the eigenvalues (taking multiplicity into account) of the operator A 17. Prove that the polynomial p(X) = II k= ,(k - X) (respectively, - p(k)) definitizes the operator A (I. Iokhvidov and M. Krein [XV] ).
4
Prove that if A is a definitizable J-selfadjoint operator, and (X,, are all the non-real roots of the definitizing polynomial p, and (k;, a;) C a(A), then a(A)\IR = (k;, X;); ,
IX,, a;] C ap(A),
Yx,(A)
and
't,;(A)
are
subspaces
(i = 1, 2, ... , k), and k
.,_ [+] (.(A) +u1'X;(A))
"Y"
i=1
and A :.JY' -.,Y' is an operator definitizable in " Wwith a real spectrum (Langer [4], [5]). 5
Prove that a definitizable J-selfadjoint operator with a real spectrum has a function which is, moreover, J-spectral E unique. Further, a(E) = a+(E) U a_ (E) U s(E), and when A fl s(E) = 0 and the A are finite the operators E(0) are the strong limits of polynomials of A (Langer [4], [5] ).
6
Prove that operators A of the class K(H) or definitizable operators are bounded if and only if a(A) is a bounded set (Azizov; Langer [4], [5]).
Hint: Use the decomposition 3.(5.2) and the results of Exercises 1, 4, and 5 respectively. 7
Prove that if A E K(H) or if A is a definitizable operator, and if a(A) _ (X;)1 , then .,Y = Yx,(A) + 99,,Z(A) + ... + 99ak(A), and the lengths of the Jordan chains of the
operators A I ',,;(A) are bounded (Azizov; Langer [4], [5]). Hint: Use the expansion 3.(5.4) and the results of Exercises 1, 4, 5. 8
Prove that if d is a commutative family of bounded definitizable J-selfadjoint operators and either .dd is finite or the `corners' P+AP- I . E .v'm for all A E d, then d has the property X111 (Langer [9], [13]).
9
Prove that if A is a definitizable J-selfadjoint operator, and if D is a finitedimensional J-selfadjoint operator, then A + D is a definitizable operator (Jonas and Langer [1]).
10
Prove that a J-selfadjoint operator of the class K(H) is definitizable if and only if function Exercise is that its E (see 1) such J-spectral a(E) = a+ (E) U a_ (E) U s(E) (Azizov).
Hint: Use Exercise 15 on 3.§ and the results of Exercises 1 and 5 taking into account a decomposition similar to 3.(5.2) for a bounded J-selfadjoint operator of the class K(H). 11
Prove that the J-spectral function of a J-selfadjoint operator A E K(H) with a(A) C R is regular if and only if all the 99),(A) are non-degenerate (Azizov). Hint: Use the expansion 3.(5.4).
12
Prove that if a J-selfadjoint operator A satisfies the condition [Ax, x] # 0 when
[x, x] = 0, then there is a constant a such that either A - aI or aI - A is a J-non-negative operator with a definite kernel and A = aI + J°°m(µ - a) dE (Personen [1], M. Krein and Shmul'yan [3]). 13
Let T be a strict plus-operator acting from a J,-space .JY, into a Jz-space .IY2, and
§2 Completeness and basicity of root vectors of J-dissipative operators 219 let a(T`T) > 0. Then the operator T has a J,-module and only one (cf. M. Krein and Shmul'yan [3] ). Hint: See the proof of Theorem 1.14. 14
Prove that a J-selfadjoint operator A is definitizable if and only if there is a function f, holomorphic in the neighbourhood of its spectrum and of the point X = oo, such that f(A) is a non-negative operator (Langer [4], [5]).
15
Prove that if A is a J-selfadjoint operator of the class K(H), then there is a decomposition Ye = Y I [+].YY2, where .', C 9-A, A.Yk'i C X j (i = 1, 2), a(A I Wi) C [m, M], - oo < m < M < oo, and A I X2 is similar to a selfadjoint operator and a(A I .YP2) C (- oo, m] U [M, co) 1. Iokhvidov [11). Hint: Use the results of Exercise 1.
16
(Azizov,
Gordienko
and
Let (U,)mm be a one-parameter group of the class (Co), (i.e., U,+S = UU5, Uo = I, and U,x U,ox when t to, and for all x E .Y') consisting of J-unitary operators,
and let iA be an arbitrary operator of this group. Prove that (A` _) A E K(H) if and only if (U,)__m E K(H) and then U, = exp(itA) (Azizov, Gordienko and 1. Iokhvidov [1]). Hint: Use the results of Exercises 1, 2, 15 and also Naymark's article [3].
§2.
Completeness and basicity of a system of root vectors of J-dissipative operators
In this section, if nothing to the contrary is stated, all the operators are continuous and are defined on the whole space. The spaces are assumed, generally speaking, to be non-separable. Our references to the monograph 1
[XI] in which the spaces are assumed to be separable are valid in this respect, that either separability is not essential for the validity of the actual results or our operators here are acting in a separable space. We recall that a system of vectors (e«) is said to be complete in a space .W if .Y' = C Lin (e,.. Everywhere it is to be understood that a E A, where A is a certain set of indices. fa) = 6(a, 0) A basis (f.) of a space Y is said to be orthonormalized if
(a, 0 E A). If T is a bounded and boundedly invertible operator, then the system (Tf') will also be a basis in YY and it is called a Riesz basis. Here we shall investigate the questions of the completeness and basicity of the roof vectors (i.e., the eigenvectors and principal vectors) of J-dissipative,
and in particular, of J-selfadjoint operators of the class K(H) and of J-selfadjoint definitizable operators. We introduce the notation (I (A) = C Lin(2'x,(A) I X E ap(A)), fo(A) = C Lin(Ker(A - XI) I X E ap(A)).
(2.1) (2.2)
By definition fo(A) C ('(A) and therefore dim. %"/ (A) S dim .Yt°/eo(A). We shall first investigate the given questions in the case of J-dissipative operators of the K(H) class. We note that if A E K(H) and if X = X E ap(A), then the root lineal 27 (A) is closed and its non-degeneracy is equivalent to its
220
4 Spectral Topics and Some Applications
projectional completeness (Exercise 5 on 3.§5). Since A E K(H), there is not more than a finite number of real points X to which correpond degenerate kernels Ker(A - X T) of this operator. We shall call the set of such points, as also in the case of J-selfadjoint operators of the class K(H) (see Exercise 1 on § 1), the set of critical points and denote it by the symbol s(A ). Lemma 2.1: Let A be a bounded J-dissipative operator of the class K(H) whose non-real spectrum consists of normal points. Then non-degeneracy of f(A) and non-degeneracy of Lin (Y), (A) I X E s(A) } are equivalent. Moreover, if f (A) is non-degenerate, then it is projectionally complete. Since (see Exercise 5 on 3.§5).,,(A) when X = A can be expressed in the
form of a sum of subspaces invariant relative to A :.,,(A) = ' [+] alt:,, where dim IV,, < co, and #), is a projectionally complete subspace and '46' C Ker(A - XI), we conclude from 2.Corollaries 2.16 and 2.18 that Lin(2',,(A)I X E s(A)) is a subspace and its non-degeneracy is equivalent to its projectional completeness. Now let Lin {27), (A) I X E s(A)) be degenerate, and let 2' be its isotropic part. Then ..2 C Lin{.il'), X E s(A)), and therefore q is finite-dimensional. By virtue
of 2.Lemma 2.19 we have AY C Y. Consequently there is in Y at least one eigenvector xo of the operator A. By its choice xo [1] Lin[114,(A)I XEs(A)), and by virtue of 2.Corollaries 2.16 and 2.18 we have xo [1] 1,, (A) when µ ¢ s(A), i.e., xo j1] 6(A). Since xo E f (A), we see that degeneracy of Lin (2'x(A )l X E s(A)) implies the degeneracy of 6(A). Conversely, let ,"(A) be degenerate, and let Y, be its isotropic part. Then AY, C 2,, by virtue of 2.Lemma 2.19. We first assume, and later prove, that dim Y, < oo. If it is, then there is in 21 an eigenvector yo of the operator A : Ayo = Xoyo From 2.Corollary 2.25 and the definition of the set of critical points we conclude that X0 E s(A), and therefore Lin(2,,(A )l X E s(A)) is a degenerate subspace. We now verify that ., is finite-dimensional, and at the same time that if f (A) is non-degenerate, then it is projectionally complete. To do this it suffices to show that there is in 44'(A) a projectionally complete subspace .4" such that dim e (A) I ./V< oo. Since A E K(H), by virtue of 3.Corollary 5.18 there are in every subspace Ker(A - XI) when X = X semi-definite invariance subspaces 9a , maximal (in Ker (A - XI)), of the operator Vo. Then r+ = C Lin(1,,(A), 2,; I Im it > 0, X = X} is a subspace completely invariant relative to Vo and invariant relative to A, and Y1- = C Lin (1,, (A ), 2- I Im µ < 0, X = X } is a subspace invariant relative to Vo
and A. By virture of 2.Corollaries 2.18 and 2.22 it follows that the first of these subspaces is non-negative and the second is non-positive. Taking
3.Corollary 5.4 into account this implies the inclusions Y'+ E h'- . Let 190 = 2,+ fl Y_ . Then Y+ + 2'_ = 20 [+] .'I where .'U is a projectionally complete subspace. It remains to notice that dim "(A) .Y'+ + 2_ < co, and therefore if f (A) is degenerate, then its isotropic part is finite-dimensional, and if it is non-degenerate, then it is projectionally complete.
§2 Completeness and basicity of root vectors of J-dissipative operators 221 Remark 2.2: The condition in Lemma 2.1 about the normality of the non-real points of the spectrum of the operator A is used only in the proof of the fact that degeneracy of f(A) implies the degeneracy of Lin (9'),(A )I X E s(A)I. Therefore the remaining assertions are valid even
without this condition, and in particular, if e (A) is non-degenerate and therefore projectionally complete, then the operator Vo I 6,'(A) belongs to the class H. In fact, the subspace Y+ constructed in the proof of Lemma 2.1 has in ,,(A) the property def 9?+ < oo. From Exercise 9 on 3.§3 we conclude that Vo I 6'(A) is a bi-non-contractive operator relative to the form [ , ] I 6,(A) and it has a completely invariant subspace JO+ E .itl+ (f (A)). In combination
with 3.Theorem5.2 and the fact that Vo E H, this is sufficient to prove the validity of our assertion. Moreover, from Exercise 17 on 3.§5 it follows that if P is the J-orthoprojector on to f (A) then the operator PVo I f (A) H and it is P*JP-bi-non-contractive.
Corollary 2.3: Let A E K(H) be a completely continuous J-dissipative operator. Then non-degenerary off (A) and 'o(A) are equivalent. The operator A satisfies the conditions of Lemma 2.1. Moreover, since all X ;4 0 are normal points of the operator A, it follows by virtue of 2.Corollary 2.25 that all the T,,(A) when X = X # 0 are non-degenerate. Consequently, degeneracy of 2o(A) and degeneracy of Lin(9?x(A)I X E s(A)) are equivalent. It remains to use Lemma 2.1.
2 The following two lemmas are key results in solving questions about the completeness of systems of root vectors of J-dissipative operators.
Lemma 2.4:
Let the operator satisfy the conditions of Lemma 2.1, let f(A) be non-degenerate, and let P be the J-orthoprojector on to Then A, = PA I (A)[) E K(H) and ap(A,) = 0. That the operator A, belongs to the class K (H) follows from the fact that if A E K(H, Vo), then by virtue of Exercise 17 on 3.§5 PVoI f,`(A)'1] E H and therefore A, E K(H, PVo I f(A)I1)Now
let X= X E ap(A). Since e '(A)1'1 is a P*JP-space, and A, is a P*JP-dissipative operator, it follows from Xo = Xo that Xo E ap(C`). Let Aixo = Xoxo. By virtue of 2.Corollary 2.17 A,xo = Xoxo. On the other hand, Al = A` I e '(A)1`1, and again by virtue of 2.Corollary 2.17 Axo = Xoxo; but this contradicts the fact that xo [1] f (A) and f(A) is non-degenerate, and therefore ap(A,) fl R = 0.
Now suppose that o # Xo E ap(A, ). By virtue of 2.Theorem 1.16 Xo E ap(Al) U a,(Al). We again use the fact that A i = A` I f (A)[1). Since by
hypothesis all the non-real points of the spectrum of the operator A (and therefore, by virtue of 2.Theorem 1.16, also of A`) are normal eigenvalues, it
follows that oEap(A`), which implies in the present case the inclusion
4 Spectral Topics and Some Applications
222
Xo E kp(A), i.e., Xo is a normal eigenvalue of the operator A and therefore also
of the operator A I "(A). Consequently, there is a decomposition f (A) = 2>(A) + such that Af' C f' and Xo E p(A I e,"). Then follows from Xo E p(A I f') that Xo E apl A2), where A2 = QA I (f' [+] f (A)I`), and Q is the
.W = Y.
.(A) + (f' [+] f (A)WWW ).
Since
Xo E ap(A1)
it
projector on to 6," [+] f (A) 1' parallel to 27x0 (A ). Let A2xo = koxo. Then (A - XoI)xo E 27>(A ), i.e., xo is a root-vector of the operator A corresponding to X0 and not lying in 27ao (A)-we have obtained a contradiction Let Jl be a certain complex of properties invariant relative to a bounded projection and an equivalent renormalization of the space, and let each bounded dissipative operator having this complex of properties have a complete system of root vectors. Then every J-dissipative operator A E K(H) with a non-degenerate f (A) also has a complete system of root vectors in M. Lemma 2.5:
We suppose the opposite. Since the properties in the complex .W are invariant relative to a bounded projection, they are in particular invariant also relative to a J-orthogonal projection. It follows from Lemma 2.4 that either
Lemma 2.5 is true, or there is a bounded J-dissipative operator A E K(H) having this complex of properties and such that ap(A) = 0. We suppose the latter holds. From 3.Theorem 5.1b we have that the operator A has at least ± ± one pair of invariant subspaces 27± E . i fl h . Since ap(A) = 0, the Y± are uniformly definite subspaces, and therefore .e = 2+ + 27-. Since the operators ±A I Y± have the properties of the complex ,* and are dissipative in relation to the scalar product ± [ , ] 127± which is equivalent to the original one, it follows by the hypothesis of the theorem that
f (A I 2±) _ 2'- and therefore up(A) - 0-we have obtained a contradiction.
Lemma 2.5 enables us to transfer to the case of bounded J-dissipative operators of the class K(H) a whole series of assertions about the completeness of the system of root vectors for ordinary dissipative operators since, as a rule,
the conditions in these assertions are invariant relative to the operations of bounded projection and equivalent renormalization of the space (see, e.g., [XI] ). We shall not cite all these assertions in the main text, but we shall introduce some of them later in the form of exercises. We show by the example of Theorem 2.6 how to carry out these exercises. We also prove Theorem 2.8;
this is interesting because it cannot be generalized, not even in the case of operators of the class H (see Example 3.3 below). But first we recall some familiar terms and introduce some new ones. .02 be a completely continuous operator, and let [s, (A )J be Let A :.W'1 the set of eigenvalues of the operator (A *A ) 1/2 (taking multiplicity into account), or in other words, [ s. (A )] is the s-number of the operator A ([XI]). We say that A E .y'p if Es. (A) < oo. In particular, if p = 1, then A is called
a nuclear operator, and if p = 2, A Hilbert-Schmidt operator. We remark that even if .,Y, is non-separable, all the s-numbers except perhaps for a count-
§2 Completeness and basicity of root vectors of J-dissipative operators 223
able set are equal to zero. It will therefore be convenient to us to suppose below that sk (A) * 0 when k = 1, 2, ... , v; v < oo. Usually the s-numbers
sk(A) are numbered in the order `biggest first', and then the formula min (II AK 11
that if equivalent
1 dim K < n) holds ([XI]). Hence, it follows in particular
are the s-numbers of the operator A in another norm
there is an m > 0 such that (n = 1, 2, ...). Let lea) be an orthonormalized basis, and { fa } be a J-orthonormalized Riesz basis of a J-space W. The number sp A = Ea(Aea, ea), where A E .9'1,
to the
first
one, then
is called the trace of the operator A. As is well-known (see, e.g.,
[XI] )
sp A = paXa, where (Xa) is the set of eigenvalues of the operator A taking multiplicity into account, and therefore sp A does not depend on the choice of the orthonormalized basis and the equivalent renormalization of the space. This enables us to write the formula for the trace in the equivalent form: sp A = Z. [Afa, fa] sign [fa, fa]. All three of these formulae will be used later.
Let A E K(H) be a completely continuous J-dissipative Theorem 2.6: with a nuclear J-imaginary component A,, and let operator
n (p; AR)Ip = 0, where AR is the J-real part of the operator A, (n = 1,,2.... ) n(p;AR) is the number of numbers of the form situated in the segment [0,p]. Then e(A)=,W if and only if Yo(A) is limp
non-degenerate.
The necessity for non-degeneracy follows from Corollary 2.2. Sufficiency: From Corollary 2.2 we obtain that if !Fo(A) is non-degenerate,
then f(A) is also non-degenerate. Let P be the J-orthoprojector on to d(A)HHH, and let A, = PA I f (A)1`1. Since (see, e.g., [XI], 2.2.1), we have limp-. n(p; (A,)R)/p = 0. Therefore if f (A) * .,Y, 11
we can by virtue of Lemma 2.4 suppose without loss of generality that the original operator A is a Volterra (i.e., A E Y. and a(A) = {0)) J-dissipative operator of the class K (H) with 0 ¢ up (A) and we shall prove that the equality limp - . n (p; AR)p = 0 is impossible. Indeed, it follows from 3.Theorem 5.16 that the operator A has invariant subspaces Y ± E, it ± fl h ± . They are
non-degenerate; for otherwise it would follow from 2.Lemma 2.19 that ap(A) 3;60. Consequently the Y' are uniformly definite and therefore .N'= M'+ + Yl-. Suppose, for example, that 2+ *- (0). We consider the It is a Volterra dissipative operator relative to the operator A+ = A (definite) scalar product [ , ] +, and limp-. n(p;AR )gyp = 0. Since the scalar product [ , ] I Y'+ is equivalent on 11 '+ to the original one, it 19+.
I
follows that limo-. n(p; AR )/p = 0, where fi(p,AR) is the number of numbers of the form are the s-numbers in the interval [0, p], and the of the operator AR relative to the scalar product [ , ] 191+. Let n+ (p; AR) be the number of numbers of the form 1/X (AR) in the interval [0, p], where the
lXn (AR)) are the positive eigenvalues of the selfadjoint operator A. Since
4 Spectral Topics and Some Applications
224
0 < n+ (p; AR) < n (p; AR +), we have limp - m n+ (p; AR )Ip = 0. It follows from
[XI] 4.7.2 in this case that limp-.. n+ (p; AR )Ip = (1 fir) sp Ai . Consequently
sp A; = 0. Since Al' is a non-negative operator, sp Al = 0 implies At = 0, i.e., A = AR is a completely continuous self-adjoint operator, and therefore ap(A+) 0-we have obtained a contradiction. We arrive similary at a contradiction if Y- ;4 (B). Later we shall more than once make use of the following simple proposition.
If A is a selfadjoint operator with a spectrum having no more than a countable set of points of condensation, then fo(A) =, and Lemma 2.7:
is an orthonormalized basis composed of the eigenvectors of the in Wthere ' operator A. Since Y) ,(A) = Ker(A - XI) and Ker(A - XI) 1 Kerr(A - µl) when X ; µ, there is in fo(A) an orthonormalized basis composed of eigenvectors of the operator A. It remains to verify that fo(A)1 = 101. Suppose this is not so.
Since a(A I fo(A)1) contains no eigenvalues, and the set of points of condensation of a(A I 6,,o(A)1) can be no more than countable, there is at least
one isolated point of the spectrum of the operator A(fo(A)1. This point must, as is well-known (see, e.g., [1]), be an eigenvalue-we obtain a contradiction.
A J-dissipative operator A will be called simple if there is no subspace invariant relative to A and A` on which these operators coincide. Theorem 2.8: If an operator A with a nuclear J-imaginary component Ai is a simple 7r-dissipative operator or if a(A) has no more than a countable set
of points of condensation, then f (A) = II, if and only if the lineal Lin(3),(A)l X E s(A)) is non-degenerate and Eq Im X. = sp Ar, where X. traverses the set of all eigenvalues of the operator A taking multiplicity into account.
Since A = AR + iA1 and the non-real spectrum of the ir-selfadjoint operator AR consists of normal eigenvalues, it follows from Ar E .9', that the non-real spectrum of the operator A consists of normal eigenvalues. Therefore
all the points of condensation of a(A) lie in R. Moreover, by virtue of 2.Theorem 2.26, a(A) _ . { "a [+] lla, where A.;{ ",, C .'{ x, dim . /I'x < oo, /la is non-degenerate, and therefore and //x C Ker(A - XI) and Lin(91>,(A)I XEs(A)) is a subspace.
Suppose that f (A)=H,. We then obtain from Lemma 2.1 that Lin ('),(A )l X E s(A )) is non-degenerate. We consider the subspace 9'= C Lin (..//a I X E a(A) fl IR). This is a non-degenerate subspace: for other-
wise it would follow from 2. Lemma 2.19 that the operator A I U' has an eigenvector xo ;4 0 isotropic in U', and by virtue of the fact that 9' is 7r-orthogonal to C Lin(. a, U',,(A)j X = X,µ * µ) it would follow that xo [l] f (A)-we obtain a contradiction. We consider the decomposition
§2 Completeness and basicity of root vectors of J-dissipative operators 225
fl = Y[+] o[l]. Since AY C M and A`Y C !, it follows that A11 C 'Ill and the operator A I -1' satisfies the same conditions as A but with this difference that all the root subspaces of the operator A I YI' are finitedimensional. Since Y C "(A) we shall suppose without loss of generality that ' = (B). Since (by construction) ./Vx Rd (B) only when X E s(A ), so aa(A) fl IR
has not more than a finite number of points, and therefore the operator has not more than a countable set of eigenvalues. Consequently, just as in [XI],
1.4.1, an orthonormalized basis (ek) can be constructed in H. such that (Aek, ek) = Xk and Xk runs through the set of eigenvalues of the operator A taking multiplicity into account. Hence, [(I/2i)(A - A*)ek, ek] = Im Xk. Since
A,E.q'1, and A,-(1/2i)(A-A*)=(1/2i)(-A`+A*)=(1/2i)(- JA*J+A*) is a finite-dimensional operator, and moreover
sp(I (JA*J-A*)I sp(21 J(A*-JA*J)J) =sp(I _
(A*-JA*J))
-sp2 (JA*J-A*)l
it follows that sp
(f (JA *J - A *) I = 0, 21 (A - A *) E J,
and sp A, = sp (2i
(A -A*)
Conversely, let Lin (Yx(A )I X E s(A )) be non-degenerate and let sp A, = Z. Im Xa. Again by virtue of Lemma 2.1.E (A) is non-degenerate. Again without loss of generality we suppose that 99 = C Lin(,/tt,, I X _ ) = (B). We con-
sider the operators A' = A E(A) and A, = PA I E(A)where P is the
7r-orthoprojector on to E (A) [lI . Since the operator A' satisfies the same conditions as the operator A, we have sp A,'= Ea Im Xa. Let (fk')) and (ffz)) be wr-orthonormalized based in 6(A) and E (A )Ill respectively. Then
sp A, = Z [A,fk'), fk')]sign [fk'), fk')] + k
= Lj k
[A,f«z),
f.(2) ]
sign [f.(2), f.(2)]
a
[Alfk'),
Im Xk +
fkl)]slgn[fkl), fk')] +
a
[(A 1),f(2), f.(2)] sign
[(A1)lfaz), 1«z)]sign
[f«z), fz)l
[f(2),1(2)]
k
which, by virtue of the equality sp A, = Ek IM Xk implies the equality [(A, )rf«z), f«z)] sign [f«z), f«z)] = 0.
Suppose for definiteness that H. is a Pontryagin space with x negative squares. Then f(A) is also a Pontryagin space with x negative squares, and therefore
4 Spectral Topics and Some Applications
226
f (A )t1J is a positive subspace. Hence it follows from Z [(At )rff2) f .M] sign
0
a
that (A,), = 0, i.e., A, = Ac. Since Ai = Ac I f (A)111, we conclude from 2. Theorem 2.15 that Af(A)111 Cf(A)[1] and AI 6(A)111 =A`I 6,(A)111 = A,. If A is a simple T-dissipative operator, it follows from this that f (A)111 = (0). If, however A is not simple but a(A) has no more than a countable set of points of condensation, then the operator A,, which is self-adjoint relative to the definite scalar product [ , ] I 6(A)[11, also has this property. Therefore, by virtue of Lemma 2.7, ap(A,) 0 if r(A)111 (0); f(A)111 = (0).
but,
by Lemma 2.1, ap(A) = 0
which
implies
that
Corollary 2.9: If A E K(H) is a nuclear J-dissipative operator, then nondegeneracy of 20 (A) is equivalent to the equality "(A) =,_W.
It follows from Corollary 2.2 that if 6(A)=, , then 2o(A) is nondegenerate. Conversely, let Yo (A) be non-degenerate. Then again from Corollary 2.2 we
obtain that 6,(A) is also non-degenerate. We make use of Lemma 2.1, by virtue of which it suffices to prove that when Y ;4 (0) there are no Volterra nuclear J-dissipative operators A of the class K(H) with 0 I< ap(A ). Let us suppose the contrary. As in the proof of Theorem 2.6 we conclude from ap(A) = 0 that the operator A has a maximal uniformly definite invariant subspace 2. Suppose for definiteness that 2 E -it' (the case 2' E ,fl- is verified
similarly). We consider the operator A I Y. This is a nuclear dissipative operator relative to the form [ , ] I Y. Since sp(A I 2') = EaXa(A I 2'), we have sp(A 12)r= Ea Im X.(A I 2). By Theorem 2.8 ap(A I .') ;4 0-and we have a contradiction. 3 Before we present results about the basicity of systems of root vectors of J-selfadjoint operators, we introduce.
Definition 2.10: A basis (fa) of a J-space 0 is said to be almost J-orthonormalized if it can be presented as the union of a finite subset of vectors and a J-orthonormalized subset, these subsets being J-orthogonal to one another. Definition 2.11: A basis (fa) is called a p-basis if there is an orthonormalized basis (ea) and an operator T E .f', such that fa = (I + T)ea (a E A ). Theorem 2.12:
Let A be a continuous J-selfadjoint operator of the class
K(H), and let a(A) have no more than a countable set of points of
§2 Completeness and basicity of root vectors of J-dissipative operators 227 condensation. Then:
a) dim J '/ (A) 5 dim , lfo(A) < oo; b) fo(A) =. if and only if s(A) = 0 and Y),(A) = Ker(A - XI) when X
X;
c) f(A)= Jt if and only if Lin(Yx(A)I X E s(A)) is a non-degenerate subspace;
d) if fo(A) =,-W (respectively, 6(A) = M), then there is in .0 an almost J-orthonormalized Riesz basis composed of eigenvectors (respectively, root vectors) of the operator A; e) if --o (A) = W, then there is in' a J-orthonormalized basis composed of eigenvectors of the operator A if and only if a(A) C FR; f) the bases mentioned above can be chosen as p-bases if and only if the
operator A has an invariant subspace Y+ E J(
with an angular
operator Ky,, E 21p.
a) Since fo(A) C f(A) it follows that dim . lf(A) < dim ,lfo(A). We use 3.Corollary 5.21 and an analogue of 3.Proposition 5.14 for J-selfadjoint operators of the class K(H) and we obtain that dim f(A)/fo(A) < oo. Consequently to prove the inequality MI-6(A) < oo. dim Yelfo(A) < oo it suffices to show that dim
Let 21+ E u+ fl h+ be an invariant subspace, existing by virtue of 3.Theorem 5.16, of the operator A, let _T- = 21_ 1I, and 21o = 2+ n 21-. Since ot'l = Lin (21+, 2-) and dim , 21,' = dim 20 < oo, to prove the inequality under discussion it is sufficient to establish that 201, = .9(A 121 ), or, equivalently, that the f space ie = XolI-To coincides with f (A ), where A is the J-selfadjoint operator generated by the operator A. But the latter follows from the fact that ie = C+ [ + ] 21 21 ± I21o are uniformly definite subspaces
invariant relative to A, and the operators A/40± satisfy the conditions of Lemma 2.7. b) Let fo(A) = M. It follows immediately from 3.Formula (5.4), from the
definition of the set s(A) and the fact that 14(A) is J-orthogonal to 2',(A) when X # µ, that s(A) = 0 and 2),(A) = Ker(A - XI) when X ;d X. Conversely, let s(A) = 0 and 2),(A) = Ker(A - XI) when X ;4 X. We again
use 3.Formula (5.4) and without loss of generality we shall suppose that a(A) C R. Since s(A) = 0, all the kernels Ker(A - XI) are non-degenerate, and therefore 2a (A) = Ker(A - XI). Hence, by virtue of Lemma 2.1, ','(A) (= fo(A)) is also non-degenerate. But since dim fo(A)WWW = dim ,/fo(A) < oo and Afo(A)11J C fo(A)111, we have fo(A)[L] = (B}, i.e..W = fo(A). c) If f (A) = ', then 6,(A) is non-degenerate, and by Lemma 2.1 Lin (2a(A) I X E s(A)) is a non-degenerate subspace. Conversely, let Lin(2),(A) I X E s(A)) be a non-degenerate subspace. Again
by virtue of Lemma 2.1 6(A) is non-degenerate and ' = 6(A) [+] f (A)1, moreover Af(A)WWW C f(A)1-, AI f(A)t1 E K(H), and ap(AI f(A)WW) =0.
228
4 Spectral Topics and Some Applications
Hence it follows that f (A )f11 = L + [+] _ , where .+ are maximal (in ('(A)1'1) uniformly definite subspaces invariant relative to A. Since a(A 12"±) C a(A), the sets a(A I Y±) have no more than a countable set of points of condensation. By Lemma 2.7 aa(A I Y+ ) = 0-we obtain a contraction with the fact that ((av(A) 12+) U ap(A I Y=) = (av(A)E(A)111) = 0.
d), e). First of all we note that if an operator A E K (H) has non-real eigenvalues, then, by virtue of the neutrality of the eigenvectors corresponding
to them and 3.Formula (5.4), there is in .W no J-orthonormalized basis composed of eigenvectors of the operator A. Now let f (a) = .1 ', and , (A) _ "a [+] ll,
when X E s(A), where A4' ,, C ,",,, dim 4'X < co, ,(4 C Ker(A - XI) and ill,, is projectionally complete (see Exercise 5 on 3.§5). Since s(A) consists of a finite number of points, and by virtue of Lemma 2.1 all the 2,,(A) are non-degenerate when X E s(A ), so, taking 3.(5.4) into account, M = '1 [+] W2, where , and W2 are invariant relative to A, and Ye, = Lin (Y,, (A ), ail "a 114 ;d µ, X E s(A)J. Consequently dim Wi < oo, and a(A 1.02) C IR1 and moreover fo(A I. 2) = f(A I Y2) =.3f2. By virtue of Exercise 15 on 3.§5 the operator A I W2 E K(H). Therefore assertions d) and e) of the theorem will be proved if we show that the conditions a(A) C 91 and fo(A) =.e imply the existence in
.w' of a J-orthonormalized Riesz basis composed of eigenvectors of the operator A. We verify this.
Let (Y+, Y-) be a maximal dual pair invariant relative to A, and and 2a = Ker(A - XI) n Y±. It can be verified, just ,
let Y± E h ±
in the proof of assertion a), that fo(A I Y+) = 2'±. Since Ker(A - XI) [ 1 ] Ker(A - µl) when X # µ, and Y + E h ± , it follows that among the subspaces Y,; for different X there are only a finite number of degenerate ones. Let these be 2',;,, 2, ... , Y +K and let 9a = Yo,> + Y;,,,,, as
where'o,a, is the isotropic part of 9', , i.e., Yo,,,, = 2X fl .a;, and the subspace a, is definite and completes 2'o,x, into Ya . Since Lin (Yo,,,, } i is a finitedimensional subspace, it follows that dim .-W/f, (A) < co, where 61 (A) = C Lin (Y, , T,-, Y 1 ,,,, 21;>,, I µ ;4 Xi, i = 1, 2, ... , n ) is a non-degenerate subspace relative to A, and moreover the maximal (in f, (A)) uniformly definite subspaces 6'1± ( A ) = C Lin (2µ , Y;,,, I µ ;4 X1, i = 1, 2, ... , n) are invariant relative to A. We now note that the operators A I f i (A) satisfy the conditions of Lemma 2.7 relative to the scalar products ± [ , ] I f i (A) respectively,
and dim fl(A)I1] < oo, and moreover
fo(A I f,(A)I') and the
kernels Ker(A I 6'1(A)1'1 - XI) of the operators (A I e', (A) 1] - XI) are non-
degenerate. Consequently in each of the subspaces f; (A ), f,- (A ), and 91,, (A I f I (A ) [1) there are J-orthonormalized Riesz bases composed of eigen-
vectors of the operator A. The union of these bases will then be the required basis.
f) Before proving this assertion we point out that, if two bases differ on a
§2 Completeness and basicity of root vectors of J-dissipative operators 229 finite number of elements and if one of the bases is a p-basis, then the second one will also be a p-basis. In proving assertions d) and e) it was essentially established that if F (A) (respectively go(A )) coincides with 'W, then there is in .1Y an almost J-orthonormalized basis composed of root vectors (respectively,
eigenvectors) of the operator A, and all these vectors, with the exception,
perhaps, of a finite number, are characteristic vectors for A and lie in pre-assigned invariant subspaces Y+ E ff + n h ± of the operator A. Let this operator have the invariant subspace 2'+ E af+ fl h+ with the angular operator KY-, E .gyp, let (fc ) be the J-orthonormalized part of a Riesz basis
composed of root vectors (or eigenvectors) of the operator A, and let ( f, } C 2 + , where T_ = 2+ ('I. Since the system (f4 } U (fq } differs from a
basis in J on a finite number of elements, it can be constructed into a J-orthonormalized basis in ' which will differ from the original one on a finite number of elements. Therefore without loss of generality we shall suppose that ( and we shall prove that it is a U If. +,) is a J-orthonormalized basis in p-basis. Since (fa } U (ff } is a basis, we have 2 + = C Lin (ff }, and therefore 11 Ksr+ 11 < 1. From 2.Formula (5.3) we construct the operator U(Kv-,), which is positive, J-unitary, and such that U(Kv-+) - I E 19p. Since (f4 } U (ff } is a J-orthonormalized basis, (e.' } U ( e,-), where e« = U-1(Ky-.) fa , also is a J-orthonormalized basis. But since e, by construction lie in .jy + , it follows that (e,+) U (e.- } is an orthonormalized basis. It remains to notice that the inclusion U(Kv-,) - I E Yp implies the inclusion U-1(Kr.) - I E .9p, and therefore U { ff } is a p-basis.
Conversely, suppose there is in . an almost J-orthonormalized basis composed of root vectors (or eigenvectors) of A and that it is a p-basis. We shall prove that there is an invariant subspace Y+ E af+ fl h+ of the operator A with an angular operator K. E gyp. To do this we separate from the basis the J-orthonormalized system (ff } U if-,,) composed of the definite eigenvectors of the operator A, with (f+ } C j P± + . We form the subspaces C Lin (ff ). They are uniformly definite, invariant relative to A, and such that def Y') < co. By 3.Theorem 4.14 the dual pair (Y+", 9 1)) can be extended into a maximal dual pair (Y+, Y_) invariant relative to A. It is clear from the construction that _T+ E lf+ fl h+. We verify that E J'p. Since dim +/ 1) < co, we can suppose without loss of generality that .?'" = Y+, i.e., (f« } U is a J-othonormalized p-basis. Let f, = (I+ T)g±, where gq } is an orthonormalized basis in , T E Jp, and eq = U-' (Kv--) f, is an + . From 2.Formula orthonormalized basis in M composed of vectors ea E . (5.3) it can be seen that Ky-. E SPp if and only if U(K1-+) - T E 9'p. We
establish this lost result. Since e,± = U-'(Ky-.)(I+ T)ga , it follows that U-' (K,,.)(1 + T) = V is a unitary operator, and therefore I+ T= V(V*U(K,,-) V) is the polar decomposition of the operator I+ T. Hence, (I+ T*) (I+ T) = V*U2 (K,-_) V, and therefore U2 (K,--) - E .f p, which implies the inclusion (U(K,,.) + I)-' (U2 (K,-_) - 1) = U(K, -) - I E SP,-
0
230
4 Spectral Topics and Some Applications
Remark 2.13: Since completely continuous J-selfadjoint operators of the class K(H) and all a-selfadjoint operators with a spectrum having no more
than a countable set of points of condensation satisfy the conditions of Theorem 2.12, the assertions in it also hold for them. Moreover, for completely continuous operators it is possible, by Corollary 2.3, to write 1o(A) instead of Lin[Y),(A) I X E s(A)] in the formulation of Theorem 2.12. 4 In conclusion we investigate the question of the completeness of the system of root vectors of definitizable J-selfadjoint operators.
Lemma 2.14: Let A be a bounded J-non-negative operator having a spectrum with no more than a countable set of points of condensation. Then the equality '(A) =.Ye is equivalent to non-degeneracy of ?o(A).
Since 2'0(A) [1] 2,,(A) when µ ;e 0, it is clear that non-degeneracy of (A) (= i) implies non-degeneracy of Yo(A ). Conversely, let .?o (A) be non-degenerate. Then 6(A) is also nondegenerate. For, if ','(A) is degenerate and if xo E e (A) fl t (A)I1I, then [Axo, xo] = 0. Consequently xo E Ker A, which implies the degeneracy of
Yo(A).
Let us assume that 4"(A) ;d °, i.e.,
(0). We consider the
operator A' = A I e(A)WWW. This operator has the following properties: it is
positive relative to the G-metric [ , J I 6'(A)111; aa(A' I = 0; and a(A') consists of not more than a countable set of points. The first two of these assertions are trivial, and so we verify the third. To do this we note first that if X E p(A), then X E p(A'). From 2.Corollary 3.25 and Exercise 7 on §1 it follows that if X0(0) is an isolated point of the spectrum of the operator A, then the
range of values of the operator A - XOI is closed, Ker(A - XoI) is projectionally complete, and Ye = Ker(A - XoI) [+] Ker(A - XOI) W1l, and moreover Xo E p(A I Ker(A - XoI)W1). Hence we conclude that if Xo(#0) is an isolated
point in a(A), then XoEp(A'). By hypothesis v(A) has no more than a countable set of points of condensation and therefore a(A') consists of not more than a countable set of points. Let µo be an isolated point of the spectrum of the operator A', and let -y,, be
a circle of sufficiently small radius with centre at the point µo. Then dX is a G-orthogonal projector on to the 'G,,-space' Ae,,, = P,04" (A) [11 invariant relative to A', and a(A' I ,,0) = [µo) . The operator Aµ0 = A' W . is G,,,, positive, and µo 0 ap(A,,0). By 3.Lemma 3.16 we extend Aµ, into a f-non-negative operator Aµ, which has 14o as the only point of its spectrum. In accordance with Exercise 7 on § 1 Aµ0 = t ol,,0 when µo ;e 0 and A20 = 0 when µo = 0, i.e., µo E op(A')-but µp(A') = 0, so we have obtained a contradiction.
Let A be a bounded J-selfadjoint operator, defmitizable by the polynomial p(X) with the roots [X;) i and let o(A) have no more than a Theorem 2.15:
§2 Completeness and basicity of root vectors of J-dissipative operators 231
countable set of points of condensation. Then the equality ','(A) =,w is equivalent to non-degeneracy of Lin
Since f(A) = f (p(A)) and 9?o(p(A)) = Lin(2),,(A));, it is sufficient to use Lemma 2.14.
Remark 2.16: Since for a J-non-negative operator A the equality 9'),(A) = Ker(A - XI) holds for all X ;4 0, so, on replacing in the formulations of Lemma 2.14 and Theorem 2.15 the root subspaces by the corresponding
kernels Ker(A - XI) of the operator A, we obtain a criterion for the coincidence of fo(A) with W. In contrast to the case of operators of the class k(H) completeness of the system of root vectors or even of the eigenvectors of a J-non-negative operator
does not guarantee the existence of a basis composed of such vectors. However, the following theorem holds. Theorem 2.17:
Let a bounded J-non-negative operator A satisfy the condi-
tions of Lemma 2.14. Then Anx = EcX « [x, fa]fa sign X., where n > 2, (Xa) C ap(A), X. 3e- 0, and (fa) is a J-orthonormalized system composed of the eigenvectors of operator A: Afa = Xa fa; if moreover A is a J-positive operator, then Ax = Z.X. [x, fa] fa sign X. (here the series converge with respect to the norm of the space W). In accordance with Theorem 1.5 Ax = Sx + 17-P dE,x. From Lemma 2.7 we conclude that
E-x_-aO, [',fa)fa, a>a where (fa) is the J-orthonormalized system of eigenvectors of the operator A corresponding to the eigenvalues (Xa) : Af« = Xafa, Xa # 0. Consequently I - Ea=
17.P dE,x = EcXc [xifc] fa sign X,,, where the series converges with respect to the norm of ,' (we notice at once that if A is a J-positive operator, then S = 0 and therefore Ax = > c Xc [x, fa] f« sign X.\). We use Theorem 1.5:
Anx= An-`IS+ _
Xn. [x, fa ] f c Y
P
dE)x= An-1
sign X.
P dEx
when n = 2, 3, ....
Exercises and problems I
Let A be a completely continuous J-dissipative operator of the class K(H), let AjE %'j and limn-,, nsn (A) = 0. Prove that f(A) = .h" if and only if 'o(A) is non-degenerate (Azizov [4], [8], Azizov and Usvyatsova [2]).
4 Spectral Topics and Some Applications
232
Hint: Use Lemma 2.4 and the definite analogue of this assertion ([XI], Theorem 4.4.2). 2
when n - Co. Prove Let A E v'. fl K(H), OA = a1p, p > 1, and o(n that ('(A) = .h' if and only if'o(A) is non-degenerate [Azizov [4], [8], Azizov and Usvyatsova [2] ).
Hint: Use Exercise 6 on 2.§2, Lemma 2.4 and the definite analogue of this assertion ([XI], Theorem 5.6.1). 3
Let A E y'- fl K(H), OA = jr1p, p > 1, and for some a for the operator B = [e'°`A]r let s. (B) = o(n-'/ P) when n - oo. Prove that the equality f (A) =.1 is equivalent to the non-degeneracy of S'o(A) (Azizov [4], [8], Azizov and Usvyatsova [2]). Hint: The same as for Exercise 2 except that [XI] 5.6.1 is replaced by [XI] 5.6.2.
4
Let
5
Prove that if in a Pontryagin G`)-space Y there is at least one almost
be a Pontryagin G(')-space (i.e., 0 E p(G('))). Prove that there is in .W' at least one `almost G(''-orthonormalized p-basis' if and only if I G(') IE yP (Azizov and Kuznetsova [1]). G(')-orthonormalized p-basis, then any other almost G(')-orthonormalized basis is also a p-basis. In particular, if G(')2 = I, then any G(')-orthonormalized basis is a p-basis for any p > 0 (Azizov and Kuznetsova [1]).
6
Give an example of a J-non-negative bounded operator B ¢ 1. such that B2 E J'P (0 < p < oo). (I. Iokhvidov [8], Azizov and Shlyakman [1]). Hint: In 2 Example 3.36 put B, E .PP and take as B2 the linear homeomorphism mapping Y, on the .Y 2, B, > 0, B2 > 0-
7
Prove that if A is a bounded J-selfadjoint operator definitizable by the polynomial p(X) with the roots (k;);, then the following assertions are equivalent: a) C\(X1) 1 CP(A); b) there is an integer q > 0 such that AQp(A) E .f'm; c) Ap(A) E ,1 ,. (V. Shtraus [3]; Azizov and Shlyakman [1]).
8
Let A E .1 be a GI')-selfadjoint operator in W (0 5 x < oo), and let 0 E a,(A). Prove that then 411(A *) = .fit') and that there is in .W(') a Riesz basis relative to the norm I x II, =III GI 112X II composed of root vectors of the operator A ([31).
9
Let A E .y'm be a G-selfadjoint operator in W, let 0 E a, (A), and let at least one of the sets (- oo, 0) fl a(GA) or a(GA) fl (0, oo) consist of a finite number of normal
eigenvalues. Prove that then rr (A ") = W and that there is in # a Riesz basis relative to the norm II x II, =III GA I "2x II composed of root vectors of the operator A Q III] ).
Hint: Relative to the indefinite form [x, y], _ [Ax, y] the space .# is a G'-space with G(') = GA. Use the result of Exercise 8. 10
11
Let ,Y be a K")-space, with 0 E p (Wt' ), let J ' I Ker W(') be a Pontryagin space, and let A be a completely continuous W(')-selfadjoint operator. Prove that there is in .# a Riesz basis composed of root vectors of the operator A if and only if /'o(A) n C Lin(21a(A)I X ;d 0) = (B) (Azizov [7]).
Prove that if A E .y',0 is a r-selfadjoint operator and f (A) * IL, then it is impossible to choose in the subspace f (A) a Riesz basis composed of root vectors of the operator A (Azizov [7] ). Hint: Prove that the inequality OF (A) ;d n, implies the inequality V'o(A) n C Lin(.V',,(A) I X ;d 0) ;d [0), and use the result of Exercise 10.
§3 Examples and applications §3
233
Examples and applications
In this section we give some examples showing the impossibility of weakening the conditions in some of the theorems given in §2, and also 1
examples showing some applications of the results in the preceding section. For our first purpose we need the following. Theorem 3.1: Let A = (AR + iAj) be a continuous operator acting in a Pontryagin space II,,, and let AR and A, be a-non-negative operators. Then non-degeneracy of ?o(A) is equivalent to the inclusion Ker A n:3?A C wA.
Let Yo(A) be non-degenerate. In accordance with 2.Theorem 2.26 Yo (A) = A 'o [+] , ifo, where /{ "o is finite-dimensional and invariant relative to A, and -Ito C Ker A is non-degenerate. Consequently A"o is a non-degenerate subspace. By construction (see the proof of 2.Theorem 2.26) the kernel of the operator A, = A I .4'0 is neutral and it is the isotropic part of the kernel of the
operator A. Since A and A, are a-dissipative operators, it follows from 2.Corollary 2.17 that Ker A = Ker A`, and Ker A, = Ker A'1. This implies the equality Ker A fl 3A = Ker A fl (Ker A`) [11 = Ker A fl (Ker A) [11 = Ker A, = Ker A, n RA,.
Since RA, C RA, it follows that Ker A n 4A C A. Conversely, let Ker A fl.A C w A. If xo E Yo(A) n 2o(A )[1), then [Axo, xo] = 0. Using the fact that AR and AT are ir-non-negative we obtain that xo E Ker A. Consequently, since xo [1] £o(A ), we have xo [1] Ker A(= Ker A`), and therefore xo EJIA, i.e., xo E Ker A n 4A. By hypothesis Ker A n 4A C RA, and therefore there is a vector yo such that xo = Ayo. The vector yo E '0(A) and therefore 0 = [xo, yo] = [Ayo, yo]
Again using the fact that AR and AI are 7r-non-negative, we obtain that (Ayo = )xo = 0, i.e., 91o(A) is non-degenerate.
The theorem just proved enables us easily to construct examples demonstrating that in the theorems about completeness (see §2) the condition of non-degeneracy of Lin [Yx(A) I X E s(A )] does not follow from any of the other conditions. We give one such example, and leave the reader to construct others. Example 3.2: (cf. Corollary 2.9 and Exercise 3 on §2 when p > 2). Let B be a nuclear operator in an infinite-dimensional space W, with Ker B ;4 (0), and let
the operators i (B+ B*) and (I/2i)(B- B*) be non-negative. In Ker B and B I RB we fix on vectors xo (11 xo
= 1) and yo (11 Yo I I = 1) respectively. Since
B is a dissipative operator, xo 1 yo. This in turn implies that the operator
J: J(«xo+ayo)=axo+ayo,
JI(Lin(xo,YoI)` = -I
234
4 Spectral Topics and Some Applications
is selfadjoint and unitary. By means of this operator we introduce the form [x, y] = (Jx, y), turning the space ,Y into a Pontryagin space with one positive
square. The operator A = JB is nuclear, and AR = J[(B+ B*)/2] and Aj = J[(B - B*)/2i], and so AR and A, are ir-non-negative operators. Since Ker A = Ker B and yo 1 Ker B, so xo [ 1 ] Ker A, i.e., xo is an isotropic vector in Ker A, and therefore xo E Ker A n A. But xo 0 ,3 A, for otherwise yo would be in ?B. Consequently, by Theorem 3.1, Yo(A) is degenerate. 2 The following example shows that in Theorem 2.8 the condition that the operator A be 7r-dissipative cannot be replaced by the more general condition that A is a J-dissipative operator of the class H.
Example 3.3: A Volterra J-dissipative operator A of the class H with a nuclear J-imaginary component and sp AI = 0. Let W = W' O+ .,Y- be a J-space, where W ± L2(0, 1). We put A;i = 0 when i P6 j,
A = II AjiMI ?i= 1,
All = -A22=2i I
r
ds
As is well-known (see, e.g., [XI], 4.7.4) A11 is a Volterra disipative operator
and (112i)(A11 - Ail) is a one-dimensional operator. So A is a Volterra J-dissipative operator with a two-dimensional J-imaginary component AI and sp AI = sp(A 11)i + sp(A22 )J = 0. It remains to verify that A E H. To do this it
± are its only maximal semi-definite invariant suffices to establish that subspaces. In the present case the latter is equivalent to the fact that for any it follows from BA 11 = A22B that B=O (see 3. Lemma 2.2). Let 2iBJ0 ' f(s) ds= -2iSo (Bf)(s) ds. Differentiating both sides of this
bounded operator B
equality with respect to t we obtain Bf = - Bf, i.e., Bf = 0 and therefore B = 0.
3 We recall that a function K(s, t) of two variables defined on a square [a, b] x [a, b] is called a Hermitian non-negative kernel if for any finite set of points (t1) i of [a, b] and for any complex numbers (;) i the sum E"J_,K(t;, t;);; is non negative, and it is called a Hermitian positive kernel if Ej",j= jK(tj, t;)E; ; = 0 if and only if , = 0, i = 1, 2, ... , n (for a more general definition see 4.§3.11 below). The function K(s, t) is called a dissipative kernel if (1/2i)(K(s, t) - K(t, s)) is a Hermitian non-negative kernel, and a strictly dissipative kernel if (1/2i)(K(s, t) - K(t, s)) is a Hermitian positive kernel. Let 9 = L2(a, b) (cf. Exercise 5 on 1.§2), i.e., the space of all w-measurable
functions sp such that lab yp(t)I2 dw(t)I < oo, where w is a function either non-decreasing or non-increasing on [a, b], and the scalar product (,p, >G) is given, up to sign, by the relation (,p, >G) = Jo p(t)>G(t) dw(t). Moreover, we I
§3 Examples and applications regard p =
235
if
Jb I P(t)-0(t)12 dw(t) = 0 a
Let a(t) be a function of bounded variation on [a, b]. In Exercises 5 and 6 on 1.§2 it was proved that the space L2, (a, b), where w(t) = WI (t) + W2 (t), and
a(t) = wi (t) - u)2 (t) is the canonical representation of a(t) in the form of the difference of two non-decreasing functions, and the space is provided with the do(t) is a J-space. In particular, if wi(t) is a piecewiseconstant function with x points of growth, then b) is a Pontryagin space with x positive squares (see Exercise 7 on 3.§9). In this and the following paragraphs we apply the results obtained in §2 to the investigation of integral operators A = Jb, K(s, t) do(t). form [gyp, '] = JQ
3.4: Let K(s, t) be a dissipative kernel continuous on [a, b] x [a, b], let A and A, be integral operators defined by the relations
Theorem
A = L K(s, t) da(t),
A, =
K(s, t) dt, J aa
a
let o(t) = WI M - w2(t), and let w, (t) be a piecewise-constant function with x
points of growth. Then if the system of root vectors of the operator A, corresponding to its non-zero eigenvalues is complete in C[a, b], then the system of root vectors of the operator A corresponding to its non-zero eigenvalues is complete in b). If in addition a(t) has no intervals of constancy, then the system mentioned of root vectors of the operator A is complete in C [a, b].
It follows from the definition of the operator A that it is a nuclear 7r-dissipative operator. Therefore by virtue of Corollary 2.9 the root vectors of the operator A corresponding to its non-zero eigenvalues will be complete in L2,(a, b) if and only if 0 ¢ ap(A). We prove that 0 0 ap(A). Let cpo E Ker A. Then po E Ker A`, i.e., JaK(s, t)(po(s) do(s) = 0. Let 1(t), ("2(t), ... be the root vectors of the operator A, corresponding to the non-zero eigenvalues and forming a system complete in C[a, b]. It follows from the definition of root
vectors that ;(t) E MA,, i.e., there are functions ti(t) E C[a, b] such that (t) = (AIE;)(t) (i = 1, 2, ...). Then b
[0o, ]'i] =
b
po(s)(AjEj)(s) du(s)
soo(s)(";(s) du(s) = J a
J a
b
a
b
'PO (s) J
K(s, t)E;(t) dt da(s) a
b
b
;(t) a
K(s, t),po(s) du(s) dt = 0. a
Since C[a, b] is denose in
b), we have po = B, i.e., 0 ¢ ap(A).
4 Spectral Topics and Some Applications
236
Now let a(t) have no intervals of constancy, and let -q,,'92, ... be a system, complete in y (a, b), of root vectors of the operator A. Since q j E A, the ?7i are continuous functions, and in b) there are functions > i such that n; = Ai,&; (i = 1, 2, ...). Let 4' be a continuous linear functional on C[a, b] such
that 4'(,q;) = 0 (i = 1, 2, ...). By virtue of Riez's theorem on the integral representation of a linear functional on C[a, b] there is a complex function of bounded variation CD such that (AOi)(s) dw(s)
4)(AO;) = J b a
= r b rb
K(s, t)O;(t) da(t) d(5(s) a
a
= J ba Oi(t)
a a
(i = 1, 2, ...)
K(s, t) &Z(s) du(t)
b), and the function w = wl + wz like a has no intervals JoK(s, t) d(,)(s) = 0 for all t E [a, b]. But then Ja ;(t) Ja K(s, t) d(:w(s) dt = 0 (i = 1, 2, ...). Since [ j';} is complete
Since if (A) = of constancy,
in C[a, b] we conclude that 4i = 0. This is equivalent to the completeness of (,t;} in C[a, b]. 4
Now let K(s, t) be a Hermitian positive kernel, and let a(t) be an
arbitrary function of bounded variation on [a, b]. We bring into consideration the iterated kernels K(")(s, t) = K(s, t), K(n) (s, t) = fab K(n-,) (s, l)K(l, t) da(l) (n = 2, 3, ... J
Moreover, we assume that the kernel K(s, t) generates a bounded operator A = JQK(s, t) da(t) having no more than a countable set of points of condensation of the spectrum. For example, if I K(s, t) 15 c < oo and if the function K(s, t) is continuous in each variable when the other is fixed, then (see, e.g., I. Iokhvidov and Ektov [1], [2]) A is a completely continuous operator and therefore has a single point of condensation of the spectrum. Let 71, denote the eigenvectors of the operator A corresponding to X. Re 0. In accordance with Theorem 2.17 for every function xE L2. (a, b) we have
A"x=
Xn [x,,7.],a(t)sign X. LT
b
_ E va' ?a (t) a
X(T)17,(T) da(T)sign X.. a
The function K(")(s, t) belongs to
b) with respect to each of the
variables. Consequently b
X_
AK(s, t) a
a
X.,)a(t),ta(s)sign X.. a
§3 Examples and applications Similarly AKA") (s, t)
X.
237
(n = 2, 3, ...
a
Here the series converge in the norm of L2,(a, b). We note also that, by definition, K"") = AK("-'). So we have proved
Theorem 3.5: If a Hermitian positive kernel K(s, t) and its iterations K(") (s, t) belong to b), and if the operator A = J K(s, t) da(t) is continuous with not more than a countable set of points of condensation of the spectrum, then when n > 2 Kt"t (s, t) _
X?7a(t),la(s)sign Xa,
(3.1)
a
where (na) is a J-orthonormalized system of eigenvectors of the operator A corresponding to Xa # 0, and the series (3.1) converges in the norm of the
space L2a, b). Theorem 3.5 has been obtained as a simple consequence of Theorem 2.17. By applying additional and different methods, going beyond the scope of our
book, for an integral operator and iterations of the kernels generated by it more precise results can be obtained (see, for example, Exercises 3-5 below).
5
In S. Krein's article [1] it is shown that the problem of the oscillations of a
heavy viscous fluid in an open fixed container reduces to the study of an operator-valued function
L(X)=XG+C-'H-I,
(3.2)
where G and H are completely continuous selfadjoint operators of finite order, i.e. G E 91P, HE 9q, p, q < oo, and G > 0, and H > 0. A whole series of general non-selfadjoint boundary problems with a parameter X in the equation and in the boundary conditions reduce to the spectral analysis of a similar operator-valued function. Here we show one of the ways of analyzing an equation (3.2) based on applying Theorem 2.13. For other results in this direction see Exercises 6 and 7 below.
Definition 3.6: In equation (3.2) let G = G ` and H = H* be bounded operates acting in a Hilbert space .'. A point Xo E C is said to be a regular
point for the function L(X) if 0Ep(L(Xo)); otherwise it is a point of the spectrum of the function L(a). A vector xo is called on eigenvector of the function L(X) if there is a Xo E C (Xo is an eigenvalue) such that L(Xo)xo = 0. The vectors x,, x2.... , x," are said to be associated with the eigenvectorxo and
the set (x;)o' is called a Jordan chain if CYJL(X0)
Z j!ax j xk-l = B ,=o
(k = 0, 1, 2, ... , m).
4 Spectral Topics and Some Applications
238
Following M. Krein and Langer [2] we introduce a scheme of argumentation which reduces the spectral analysis of the function L(X) to the spectral
analysis of a certain J-selfadjoint operator. We shall suppose that G > 0. After replacement of the variables X _ - µ ' - a the function L (X) becomes the function
L,(µ)
µ(1 +µa)
[µ2(a2G+ H+ aI)+µ(2aG+ I) + G].
We put
a>inf(b>01 Fb= b2G+H+b1>0). Then
Li(µ)= - µ(1 +µa) Fa (µ2l+µBa+Ca)FQ/2, 1
where
Ba=Fa "2(2aG+I)FQ v2)> 0,
=Fa'2GFa''2
Ca
It can be verified immediately that Xo is an eigenvalue of the function L(X) if and only if µo = - (Xo + a) - ' is an eigenvalue of the function L2(µ) =,U21+ µBa + Ca, called a quadratic bundle. Moreover, (xo, xi, ... , is a Jordan chain of the function L(X) if and only if is a Jordan chain of the bundle L2(µ). (FV2xo, FcV2x,, ... ,
_ .i+ O+ e-,
We bring into consideration the J-space the J-selfadjoint operator 0 Caln
Aa =
Ve
_' and
CQi2l (3.3)
-Ba
acting in it. It is easy to see that the regular points, the spectrum, and in
particular the eigenvalues of the bundle L2(µ) and of the operator Aa coincide. Moreover, if (yo, yl, . . . , y.,.) is the Jordan chain of the bundle L2(µ) corresponding to the eigenvalue µo, then the vectors Co /2 Yo
C.1/2 Y1
ILOyo
µ0y I + Yo
Cav2
µoyn, + yin- I
form a Jordan chain of the operator Aa, and conversely if zi') zf2)
is a Jordan chain of the operator A., then yo=
I AO
z62),
(z(?)yi= 1 (zf2)- yo),..., y,,,= 1 µo µo
is a Jordan chain of the bundle L2(µ).
y,,,-I)
§3 Examples and applications
239
Thus, a one-to-one correspondence has been established between the Jordan
chains of the function L(X) and the operator A. Here we denote by symbols .4->, and -lix
KerL(X) fl ((2XG- I)Ker L(X))1
KerL(X)fl (Lin((2XG-I)xk+Gxk_,)o")1 respectively, where (xo, x,, ... , x,,,) are all possible Jordan chains of the and
bundle L (X) corresponding to the eigenvalue X, and x_ 1 = 0.
Let the function (3.2) be given, where G > 0, G E 99.,, Theorem 3.7: H = H* is a bounded operator, and the set a(H) is no more than countable. Then:
1) each of the operators (3.3) generated by the function L (X) belongs to the class H;
2) dim .'/f(Ao) < dim Y16-o(A,,) < oo; 3) fo(AQ) =.W'
if and
only
(211G11)-'IxIs211H11;
4) f (A0) = 1Y if and (211 GII)-'
only I XI s211H11;
if
.
"\ = (0)
for all
X
such
that
if
/1"a = (0)
for all
X
such
that
5) If fo(A.) = W (respectively, f (Aa) = M), then there is in W an almost J-orthonormalized Riesz basis composed of eigenvectors (respectively, root vectors) of the operator Aa. If fo(A,,) = X, then there is in W a J-orthonormalized Riesz basis composed of eigenvectors of the operator
A. if and only if the function L(X) has no non-real eigenvalues. Moreover, the bases mentioned above can be chosen as p-bases if and only if G E 9'p12-
1) Since it follows from G E Y. that Co/2 E 9'- so (cf. 3.Theorem 1.13, El 2.Remark 6.14, and 3.Theorem 2.8) the operator A has at least one maximal
non-negative invariant subspace 9+ with an angular operator K. In accordance with 3.Lemma 2.2 Kr--CQ"2K, + C"2 + B,,K'-- = 0, and since B,, )> 0, so Kv~ = - BQ'(CQ"2 + K,-CQ,2K,--) E q.. Consequently (cf. Exercise 3 on 3.§5), A E H, and moreover, Kv,. E .%'p if and only if CQ12 E 9p, which
in turn is equivalent to G belonging to the class .9'p,2 (see [XI], 3.7.3). Assertions 2) to 5) will now follow from Theorem 2.13 if we show that
the set a(Ao) is no more than countable, that the non-real eigenvalues of
the
operator
Aa
and
the
set
(t=-(X+a)-' 1(211G1-'slxl<211H11),
s(AQ)
lie
the set ./VX=(0) and in
and that = (0) are equivalent to non-degeneracy of Ker(A,, - µI) + Ker(A,, - AI) and 2 (AQ) respectively when µ = (X + a)-'. That the set v(A,,) is no more than countable follows from the fact that the operator Ao is the sum of the operator 0
0
0
- Fo '
240
4 Spectral Topics and Some Applications
and the completely continuous operator Ca z
0
- 2 aFa 1/2 GFa
- cJ"2
1"2
and the spectrum of the first of these consists of not more than a countable
set of points, since the operator H, and therefore also the operator Fa = a 2 G + aI + H, has this property (see 2.Theorem 2.11). We now verify that all the non-real points of the spectrum of the function Y'(X) are situated in the ring (2 II G II)-' S I X I S 2 II H II . For, if L (Xo)xo = 0 and Xo ;x-I o, then, solving the quadratic equation X o(Gxo, xo) - Xo(xo, xo) + (Hxo, xo) = 0,
we obtain Xo =
(xo, xo) ± ,(xo, xo)T- 4(Gxo, xo)H(xo, xo) 2(Gxo, xo)
Hence,
I x0 2- (Hxo,xo)-2(Hxo,xa). (xo,xo) (Gxo, xo)
2(Gxo, xo)
(xo, xo)
But since the discriminant of the equation is less than zero, we have (xo, xo) < 2(Hxo, xo) 2(Gxaxo)
(xo, xo)
and therefore
(xo, xo) 1 z\ I Xo I z\ 2(Hxo, xoz `2(Gxo, xo)J
(xo, xo)
Hence, we conclude that X0 is situated in the ring (2 II G II) -' S I X I S 2 II H II Consequently the non-real spectrum of the operator A coincides with points µo
of the form µo = - (Xo + a)-'.
We now show that the set a(Aa) is situated in the same ring. Let µ, = µ, E up(Aa) and let Ker(Aa - µI) be degenerate. As proved earlier this is µi ' - a) a vector equivalent to the fact that there exists in Ker L (X,) (X, xo such that the vector CaizF./Zxo
-
1
+a
F/2xo
is J-orthogonal to all vectors of the form Caiz1vzy
,+a Fvzy 1
§3 Examples and applications
241
where y E Ker L(X1), i.e., (Ca/ 2Fa"2xo, Ca/2Fo'/2y)
- (X1 + a) 1
2
(Fa12xo, Fa12y)=0.
Starting from the definitions of the operators C. and F. we obtain as an equivalent form of this equation
((X G + 2aX, G - I- H)xo, y) = 0.
(3.4)
We substitute in this equation the vector Xtxo - X1Gxo instead of the vector
Hxo and, after cancellation by X1 + a we obtain ((2X 1 G - I)xo, y = 0. In particular, when y = xo, we have X = (xo, xo)/2(Gxo, xo). But if in (3.4) we substitute instead of X i Gxo the vector X 1 xo - Hxo and instead of X 1 Gxo the
vector xo - (l/X)Hxo, and cancel out X, + a (when y = xo) we obtain that X, = 2(Hxo, xo)/(xo, xo). From this we conclude as before that X, lies in the
ring (211 GII) 'S IXI <2IIHI1. The validity of the interpretation given above of the equalities,'Ux = (B) and ..ffa = (0) is proved by entirely similar considerations.
Remark 3.8: In the proof of the equality f (A.) = , it is necessary to verify the triviality of Ji, only for those points of s(AQ) which are points of condensation of the spectrum of the operator Aa. But the latter coincide with the set of points of the form µ = - (X + a) - ' where X traverses the set of points of condensation of the spectrum of the function (3.2). If G, HE .9pe, then it
follows from the form of the operator A. that
it
has two points of
condensation of its spectrum : µ, = 0 and µ2 = - a - '. Consequently the function (3.2) also has only two points of condensation of its spectrum: X1 = 00 and X2 = 0 respectively. Both these points do not lie in the ring (211 G I I ) - ' S I X 15 2 II HII, and therefore when G, HE 9'm we always have
f(A)='.
Definition 3.9: The function L (X) = X G + (1 / X )H - I is called a strongly damped bundle if (x, x)2 > 4(Gx, x)(Hx, x) for all x ;e 0. Corollary 3.10: If under the conditions of Theorem 3.7 the function (3.2) is a strongly damped bundle, then there is in W a J-orthonormalized Riesz basis
composed of eigenvectors of the operator A, and it will be a p-basis if GE 1'p,2.
By Theorem 3.7 it suffices to verify that the function (3.2) has no non-real eigenvectors and that A'A = (0) for all X. The first follows immediately from
the fact that if X is an eigenvalue of the function (3.2) and if xo is the corresponding eigenvector, then Xo =
(xo, xo) ± .(xo, xo) - 4(Gxo, xo)(Hxo, xo) _ -o 2(Gxo, xo)
4 Spectral Topics and Some Applications
242
Let us suppose that for some Xo the subspace ,/I'>, ;e (0). Then, in particular, some xo E Ker ((2Xo G - I)xo, xo) = 0 for L (Xo), and therefore Xo = (xo, xo)/2(Gxo, xo), which in turn implies the equality (xo, xo)z =
4(Gxo, xo)(Hxo, xo), and we have a contradiction with the fact that the function (3.2) is strongly damped.
Exercises and problems 1
2
Investigate the possibility of generalizing Theorem 3.1 to the case of a Krein space.
let K(s, t) be a Hermitian positive kernel continuous on the square 0 5 s < t < b, let R(s, t) be a function bounded on this same square and continuous with respect to each variable when the other is fixed, let R(s, t) = R(s, t), and let a(t) generate a
Pontryagin space according to the form [,o, ,&J = la do(t), and let A = Jo (R(s, t) + iK(s, t)) do(t). Prove that the equalities F (A) = L2 (a, b) and JQ K(t, t) do(t) = E; Im k;, where (a;) is the set of eigenvalues of the operator A, are equivalent. (Azizov [8] ). 3
Consider on [a, b] x [a, b] a Hermitian bounded kernel K(s, t) continuous with respect to each variable when the other is fixed, and a function of bounded variation
a(t) 0 const. We shall say that the kernel K(s, 1) is a-non-negative if K(s, t)g(s)g(t) da(s) da(t) >, 0 for all g E a
Y,
a
where )/" is the lineal of all those functions from L2 (a, b) (o = Var a) which are generated by continuous functions from C[a, b]. Prove that a-non-negativity is equivalent to either of the conditions: a) J b K(s, t) dr(s) dr(t) > 0 for any (complex) function of bounded variation r on [a, b] with E, C Ea (E, and E. are the sets of points of variation of the functions r and a respectively); b) Y- k=, K(t;, tu)Eksk ,>O for all n E N, ti, t2, .. ,,, E Ea, and
(I. Iokhvidov and Ektov [1], [2]). 4
Let K(s, t) be a a-non-negative kernel satisfying the conditions of Exercise 3. Prove that the assertions `K(2 (s, t) = 0 on E. > Ea' and 'K(2 (s, t) = 0 on [a, b] x [a, b]' are equivalent (I. Iokhvidov and Ektov [1], [2]).
5
Prove that the integral equation p(s) = XJo K(s, t),p(t) da(t) with a a function of bounded variation, and a a-positive kernel, has x < oo positive characteristic points if and only if the positive variation of the function a has exactly x points of growth (M. Krein [1], I. Iokvidov and Ektov [1]).
6
Let S=S*E.y'., T=T`E.y'm, -1Ep(s),0>iao(T)andA=(I+S)T.Prove that 6 (A) = .0 and that in i' there is a Riesz basis composed of root vectors of the operator A ([XI] ); if S E .vo, then there is in .W' a p-basis composed of root vectors
of the operator A (Kopachevskiy [1]-[3]). Hint: The operator A is 7r-selfadjoint relative to the form [ It only remains to use Remark 2.13 and Exercise 4 on §2. 7
,
] _ ((I+ S)
'
,
).
Prove that if L (k) is the function (3.2), G = G * E .v'm, H = H* E .'/'. and Ker G =
Ker H= (0), then:
Remarks and bibliographical indications on Chapter 4
243
1) the system Xk(k)
{
(-ly k
j=0
j+1 xk-j(t)
1 of `special vectors', where (xo(X), ..., xk(k) ( is the Jordan chain of the function L(X) corresponding to the eigenvalues k, is complete in .JY (Askerov, S. Krein and Laptev [ 1 ] );
2) in ,Y there is a Riesz basis composed of vectors of the special form (Larionov [6], [7]); 3) If G E yo, HE 'Q, and r = max(p, q), then there is in -Y an r-basis composed vectors of the special form (Kopachevskiy [1]-[3]).
Remarks and bibliographical indications on chapter IV §1.1. The J-spectral function was introduced by M. Krein and Langer [1] for the case of IL, and later Langer [4], [5] transferred the investigation to the case of Krein spaces. The definitions in the text, given by Azizov, modify Krein and Langer's definitions in order to accommodate them to operators of the K(H) class (see Exercises 1 and 2). §1.2-1.3. Definitizable operators in H. were introduced by I. lokhvidov and
M. Krein [XV], and in a J-space by Langer [4], [5], [9]. All the results of these sections are due to him. We borrowed from Bognar [5] the elegant proof of Theorem 1.5. In Bognar [5] these are also given bibliographical references
to other proofs and, in particular, references to M. Krein and Shmul'yan's proof [3] based on considerations from the problem of moments. In connection with criteria for the regularity of the J-spectral function see Langer [4], [5), Jonas [1]-[7], Jonas and Langer [1], [2], Akopyan [1]-[3], and Spitovskiy [ 1].
§ 1.4. Definition 1.11 and Theorem 1.12 for the case of operators acting from one space into another are modifications of corresponding results of Potapov [ 1], Ginzburg [2], M. Krein and Shmul'yan [3]. On square roots of J-selfadjoint operators see Bognar [ 1] and, more fully, Bayasgalan [ 1]. §2.1. The results in this paragraph are due to Azizov. Corollary 2.3 was published in the paper by Azizov and Usvyatsova [2]. §2.2. Lemmas 2.4 and 2.5, Theorem 2.6 and Corollary 2.9 in the case A E Y. were published in the paper by Azizov and Usvyatsova [2]. The presentation in the text is due to Azizov. To him also is due Theorem 2.8, which is a transfer to IIX of a corresponding result from [XI]. Lemma 2.7 is, apparently, well-known, though we have not found its formulation in print.
§2.3. Theorem 2.12 is due to Azizov [13]. For the formulation of the problems about p-basicity he is obliged to Kopachevskiy, who first began the
study of this question for the case of indefinite spaces in connection with problems of hydrodynamics (see, e.g., Kopachevskry [ 1]- [3]). The concept of p-basicity itself was introduced by Prigorskiy [ 1]. We mention also that the
244
4 Spectral Topics and Some Applications
questions of completeness of a system of root vectors of 7r-selfadjoint operators A E .`/'- with Of aa(A) were investigated for the first time by 1. lokhvidov [2]; the existence of a Riesz basis of root vectors of such operators was proved essentially in [XI]; the criterion for completeness and basicity of these vectors without the condition 0 0 ap(A) was given by Azizov and I. Iokhvidov [1]. On other conditions for completeness and basicity and for a historical survey see [IV] and also the Exercises. §2.4. The results of this paragraph in so general a formulation are due to Azizov. In the case when the set a(A) has no more than a finite number of points of condensation, Theorem 2.15 and Remark 2.16 (even in the case of Banach spaces with an indefinite metric) were obtained earlier by Azizov and Shtraus [ 1 ] , and Theorem 2.17 when A 2 E 9 " . is due to Kuhne [ 1 ] (see 1. lokhvidov [ 8], Ektov [ 1 ] ). As above, for a historical survey and for other results we refer the reader to the survey [IV] and to the Exercises on this section.
§3.1-3.3. All the results in these paragraphs are due to Azizov. Some of them were published in Azizov's paper [8]. Theorem 3.4 for the case of a Hermitian kernel was obtained earlier by I. Iokhvidov [2]. §3.4. Theorem 3.5 in the text was proved by Azizov; for the case A E 3'm see M. Krein [1], I. Iokhvidov and Ektov [1]. §3.5. Theorem 3.7 and Corollary 3.10 were proved by Azizov [13] (cf.
Askerov, S. Krein and Laptev [1], M. Krein and Langer [2], Larionov [6], [7], Kopachevskry [1]-[3], Azizov and Usvyatsova [2]). As regards other investigations of operator bundles and application, of an indefinite metric see Langer [4], [6], [10], [12], Kostyuchenko and Orazov [ 1 ] , and others.
5 THEORY OF EXTENSIONS OF ISOMETRIC AND SYMMETRIC OPERATORS IN SPACES WITH AN INDEFINITE METRIC
In § 1 the apparatus of Potapov-Ginzburg transformations is developed, and its application to the theory of extensions is demonstrated. §2 is devoted to another approach to the theory of extensions of isometric operators in Krein spaces. In §3 generalized resolvents of J-symmetric operators are described.
§1
Potapov-Ginzburg linear-fractional transformations and extensions of operators
One of the methods allowing extensions of (Jr, J2)-isometric operators to be constructed is the application of Potapov-Ginzburg transformations or, 1
briefly, PG-transformations. Definition 1.1: Let .,Y, =X1' Q+ .YP,- and .YP2 = .02' Q+
. z be the canonical decompositions of a J,-space ., and a J2-space .02 respectively, and let 1(3 = .Wi O+ )Yz and ,Y4 = ,Yz OO i"i be a J3-space and a J4-space constructed from them with J3 = It Q+ - Iz and J4 = Iz (j - IF. A transformation w+ :.r°1 O+ .#2 -- H3 .W'4 carrying a vector (x,, x2) into a vector
245
5 Theory of Extensions of Isometric and Symmetric Operators
246
(X3, X4), where
XI=Xj +Xl
X2=X2 + X 2
,
X3=X1 +X2
,
,
X4=X2 +X1
,
,+ (i = 1, 2), is called a PG-transformation. E As well as the transformation w+ having the form
x,
w+:
1 QIY2--,IY3 Q+'4,
W
other PG-transformations can be considered: U)
:.Y1 G. 02-x,03 Q W4,
W±:
w (XI,X2)= <-X1 +X2,X2 +X1 ), W±<X1,X2)=(X2 ±X1,±XI +X2).
-W2-J4('Y3
Below we shall study the properties of the transformation w+. The properties of w- and w+ are similar, and we leave it to the reader to formulate them and prove them by the same scheme as for w+ or relying on the following obvious proposition: 1.2.
W- = w+J,
w- (X1, X2) = (w± (x1, x2))-1,
CO- = W+J,
where
J(x,y) = (-x, y)
and
((x,Y))-' = (Y, X).
We note also that the spaces W3, ,4 and i, M2 may interchange roles, and the transformation w+ can be applied to a vector <x3i x4) E. 3 ( ,W'4. In
what follows we shall cite definitions and propositions for one of these variants, remembering that they are symmetric.
1.3. The transformation w+ maps ,Y, O+ .2 one-to-one on to .3 (D and it is involutory: w+(w+<xl, x2)) = <xl, x2), and if (X1, X2),
4,
and (X3, X4) = w+<X1, X2), (Y3, Y4) = W+(Y1, Y2
then
[xl, Y1 I I - [X2, Y212 = (X3, Y3)3 - (X4, Y4)4
here ( , )3 and (
,
)4 denote the scalar products in .7(3 and
4 respec-
tively.
The vector < X3, X4) = w+ < X1, X2) is defined uniquely by the pair < xi, x2)
because projection of xi on to Wi± (i = 1, 2) is single-valued. The existence of the inverse transformation, its definition on the whole space .'Y3 O+ .7(4, and the equality w+(W+<x1, x2) are obvious. It remains to verify the identity (1.1): [x1,Y1]1- [x2,Y2]2=(x1+,Y1 )I-(X17, Y? )1-(X2+, Y2 )2+(x2 ,Yz )2
=(Xi +xz ,Y1 +yz )3-(x2+ +x1-,Y2 +Yi )4 = (x3, Y3)3 - (X4, Y4)4.
§1 Potapov-Ginzburg linear fractional transformations
247
2 We define the PG-transformation w+ on sets ./VC .1 O+ .02: (1.2)
w+(.4) = (w+<xl, x2) I <x1, x2) E.4].
From Definition 1.1 and formula (1.2) it follows immediately that 1.4 If 2' is a lineal in '1 O+ .2, then w+ (Y) is a lineal in A3 + .r4. El
Definition 1.5: Let T :.'1 -+ .02 be a linear operator. We shall say that T E T if Ker(PZ TPi I.3 'j ) = (B) , and we introduce the PG-transformation
w+ (T) :3 -'4 for such operators by the formula w+ (T) _ (Pi + PZ T)(Pi + PZ T)-
(1.3)
(the existence of the operator (Pi + Pz T)-' follows from Exercise 6 on 2.§1). Theorem 1.6:
If the operator TE J and if rr is its graph, then
w+(rr)=r,-(r) and w+(T)E T. Since 9,,-(T)= ((Pi + Pz T)x1 I xl E fir) and FT= (<xl, Tx1) I xl E cT), we have, using (1.2) and the Definition 1.1, w+(rr) = (w+<x1, Tx1) I x1 E VTJ
_ (((P; + Pz T )xl, (Pi + PZ T )xl) I X1 E CAT) _ (<(Y3, (Pi + P3 T)(Pi + Pj T)-'Y3) I Y3 = (Pi + Pz )xl, x1 E !2T) = r.+ (r).
Since w+ (T) acts from .)r3 into W4, so, in accordance with Definition 1.5 and Exercise 6 on 2.§1, w+ (T) E J if Ker(P3 + Pa w+(T)) = (0). Let
Y3 E Ker(P3 + Pa w+ (T));
Y3 = (Pi + Pj T)x1 E cA-(r), x1 E CAr.
Then
0=(P3 +P4w+(T))y3=P3(Pi +Pi T)x1+Pa(PI +P3T)x1 = xi + ...XI = x1. Consequently,
y3 = 0,
i.e.,
Ker(P3 + Pa w+ (T)) = (0),
and
therefore
w(T)E T. Corollary 1.7: The transformation (1.3) establishes a one-to-one correspondence between (J1, J2)-non-expansive operators V:.11 -.y'2 and con-
tractions W:.3 -p .4, WE T. Moreover, V is a (J1, J2)-bi-non-expansive operator (respectively, a (J, J2)-isometric operator; a (J1, J2)-semi-unitary
operator; a (J1,J2)-unitary operator) if and only if W=w+ (V) is a contraction, V w =
.
3 and R p,- wP_ = .01 (respectively, W is an isometry; W is
an isometry and gyp,- wpb =i ; W is a unitary operator and 5?p,- wp, =,01).
248
5 Theory of Extensions of Isometric and Symmetric Operators
It follows from Exercise 33 on 2.§4 in conjunction with 2. Remark 4.29 that strict minus-operators (including (J1, J2 )-non-expansive operators) VE J. It only remains to use Theorem 1.6, the Formulae (1.1) and (1.3), the
definitions of the corresponding classes of operators, and the results of Exercise 33 on 2.§4 and Exercise 7 on 2.§1. For, let us carry out, for example, the appropriate argument for the case when V is a (J1, J2)-unitary operator. It follows from Theorem 1.6 that WE J. From (1.1) we conclude that W is an isometry, and from (1.3) in combination with Exercise 33 on 2.§4 that 9 w = IY3. Similar considerations show that the domain of definition of the operator
w+(V)
(P+ +PZ V) (Pi +Pz V)-':4-.3
.4. It remains to observe that (cf. Proposition 1.2) w+ (V) _ (w+ (V)) ' = W-', and therefore W is a unitary operator. Since Pi WPz I . z = (Pi VPi IM (see Exercise 1 below), ?p,- wP_ = . i Conversely, let W be a unitary operator and let 91P,- wp, = W_-. One proves, is
as we did above for W, that the operator V is defined on '1, and V is a (J1, J2)-isometric operator, i.e., V is a (J1, J2)-semi-unitary operator. From the facts that the operator WE J is unitary and that 91P,- wP. = ,Y1 it is obtained immediately that
Ker(P2 WPi I.W )2 = [e1,
J?P, wp, = w2+,
and, using the same considerations as above for W, we obtain that the operator V -' is (J1, J2)-semi-unitary, and therefore V is a (J1, J2)-unitary operator. IY2. We recall that an 3 Again let rT be the graph of the operator T:.Yf 1 operator Tis an extension of the operator Tif and only if rT C Pr. It therefore follows from Theorem 1.6 and Corollary 1.7 that:
1.8 An operator T(E T) admits an extenson T E 1 i and only if the operator w+ (T) has an extension w' (T) E T, and then w+ (T) = w+ (T).
In particular, a (J1, J2)-non-expansive operator (a (J1, J2)-isometric operator) V -admits a (J1, J2)-non-expansive extension (a(J1, J2)-isometric
extension) V if and only if the contraction (isometry) w+(V) admits a contractive (isometric) extension w+(V)E .1, and then j)=w+(V). We now pass on to the description of special extensions of (J1, J2)isometric operators. Definition 1.9:
.2
An extension T of a (J1, J2)-isometric operator V:.'Y1 is called its (J1, J2)-bi-extension if Fv is isotropic in rT relative to the form (x1, x2),
From Definition 1.9 and the fact that graphs of (J1, J2)-non-expansive operators and (J1, J2)-non-contractive operators are semi-definite lineals L]
§1 Potapov-Ginzburg linear fractional transformations
249
relative to the form 2.(4.11), and the graph of a (J1, J2)-isometric operator is a neutral lineal, we obtain, taking 1.Proposition 1.17 into accounts. 1.10 (J1, J2)-non-expansive, (J1, J2)-non-contractive, and, in particular, (J1, J2)-isometric extensions of a (J1, J2 )-isometric operator are (J1, J2)-biextensions of it. The following proposition also follows directly from the Definition 1.9: 1.11
Let T be an extension of a (J1, J2)-isometric operator V If T is a
(J1, J2)-bi-extension of it, then T(r!r n vv,') C j? 1 Conversely, if CA v is a
projectionally complete subspace and T(VT n cV)) C *fr1, then T is a (J1, J2)-bi-extension of the operator V From the Definition 1.9, from 2.Formula (4.14), and from the fact that
the graphs Fv= (<x, Vx) I XE tv) and Fv- = (( V-1y, y) yE liv-t) of the operators V and V -' respectively, regarded as lineals in W1 + M2, coincide, it follows that the following proposition holds: 1.12. Let V be a (J1, J2)-isometric operator with Ker V = (0), and let T be an extension densely defined in .r1 of the operator V Then the operator T is a (J1, J2)-bi-extension of the operator V if and only if V -' C T` and therefore
the operator T` is a (J2, J1)-bi-extension of the operator V -'. The following theorem establishes a connection between (J1, J2)-biextensions of a (J1, J2)-isometric operator and bi-extensions of its PGtransformation. Theorem 1.13: An extension T of a (J1, J2)-isometric operator V is a (J1, J2)-bi-extension of V if and only if w+ (P v) (C .3 (B 04) is isotropic in w+ (Pr) (C.3 O+ 4) relative to the form (X3, y3)3 - (x4, Y4)4. In particular, if T E T, then T is a (J1, J2)-bi-extension of a (J1, J2)- isometric operator V if and only if the operator w+ (T) is a bi-extension of the isometric operator We use the Definition 1.9 and we also note that the left-hand side of the
formula (1.1) coincides with the (- J1,)-metric. Therefore (x1, x2) is an isotropic vector in P r if and only if w+ (x1, x2) = (X3, x4) is an isotropic vector in W' (FT) relative to the form (X3, y3)3 - (X4, y4)4. From this the first assertion of Theorem 1.13 is obtained. The second assertion is a direct consequence of
the first, taking Theorem 1.6 into account.
Corollary 1.7, Proposition 1.8, and Theorem 1.13 show us the way to describe special extensions of (J1, J2 )-isometric operators (see Exercises 5-11). It is well known that in extensions of symmetric operators in a Hilbert space the method of Cayley-Neyman transforms is used, reducing the given problem to the problem of extending isometric operators. In J-spaces such a method is not always applicable, since (cf. 2.Example 3.36) it is not excluded that ap(A) = C. However, in 2.Proposition 2.3 we mentioned another method 4
5 Theory of Extensions of Isometric and Symmetric Operators
250
of extending J-dispative, and therefore also J-symmetric, operators: JA - JA JJA. Finally, in the preceding paragraphs a method of
A
extension by means of PG-transforms has been indicated.
It turns out that the three transformations-the Cayley-Newman transform, multiplication by J, and the PG-transformation-are closely connected.
Theorem 1.14: Let . ' Wbe a J-space and Or =,-W @.,Y be the space of graphs. Then K; 0 [ J, IJ 0 K; ' (x, y) = w+ (x, y) for all (x, y) E Or, here
[J,1]0(u,v) = (Ju,v) ((u,v) E fir).
D K;°[J,I]
°K;1
(x, y)=K;°[J,I] °
x+y x+y 2i
'
2i
-K`Jx+Jy x+y -(x++y-,x +y+)=w+(x,y). 2i
21
From this follows directly.
' be a linear operator: Then: Corollary 1.15: Let T: ,Y( a) TE T, 1 up(T) iOvv(Ki'(T)J); b) TE T, l ap(w+T) = 1 ¢ap(T):
c) 10up(T), i0up (K;'(T)J)- TE T. If the premise of at least one of the conditions a), b), or b) holds, then K;(K; '(T)J) = w+ (T).
§ Exercises and problems 1
Let T:411 ?Pz TP,- _
i Q1 ) .2
z O+ z ), TE .1T, Prove that then V,,-(T) = IY3 (_ i (B -Y2- ), and if T = II T%i ?i=z is the matrix representation of the operator T, then W+(T)1(w+(T))ijl z
w+(T):J''3-y-W4 (= z
D.W
)
has the matrix representation
j=i, where (w+(T))u = T11- T12Tz2'T21, (w+(T))12 = T12Tzz', (w + (T))21 = Tzz' Tz,, (w + (T))22 = Ti-21 (Shmul'yan [5] ). 2
Prove that the PG-transformation w+ establishes a one-to-one correspondence between all (J,, J2)-non-expansive linear relations .4 " C /', O+ .W'2 (i.e., [x2, x2] < [x,, x,] if (x,, xz) E./I") and all contractions V:.W3 -.YP4. Investigate what is the pre-image of isometries, of semi-unitary and of unitary operators (Shmul'yan [5], Ritsner [4]). Hint: Use Formula (1.1).
3
Construct an example of a (J,, J2)-bi-extension of a (J,, J2)-isometric operator which is neither a (J,, J2)-non-expansive operator nor a (J,, J2)-non-contractive operator. Hint: Use Proposition 1.11.
4
Derive formulae connecting the Cayley-Neyman transform Kr for an arbitrary and the Potapov-Ginzburg transformations .., w± (Azizov, E. lokhvidov, and I. lokhvidov [1]. Hint. Use the scheme of the proof of Theorem 1.14.
§1 Potapov-Ginzburg linear fractional transformations 5
6
251
Prove by means of PG-transformations that a continuous, continuously invertible, closed (J1, J2)-isometric operator with finite deficiency-numbers p = -dim 9v 11 dim ,Wv11 admits a (J1, J2)-unitary extension if and only if In 1111 = In i'l 11 (cf. E. Iokhvidov [4], [5] ). Let V be a bounded (J,, J2 )-isometric operator with Vv= 1 v, p = dim 'J
q=dim
X11,
.
r=dim Ker V, m=dim
n=dim
Then
p + q = in + n + r (Azizov, E. Iokhvidov, I. Iokhvidov [ 1 ]; cf. E. Iokhvidov). 7
Give examples of operators for which the numbers p, q, r, in, n in the equality in Example 6 take all the possible values admissible by this equality (Azizov; cf. E. Iokhvidov [4] ).
Hint: Consider a J-space
.
= .YP+ 0,,Y- (where .YP+ = .YY- is an infinite-
dimensional space), and a semi-unitary operator, acting in .h°, U= 11 U;ijj?i=j, where Uii = -K*U2i, KE.X+; U22 = KU12; U21 = S21(S2*IKK*S2i +
1+)-in
is a operator .Yl'+ into semi-unitary carrying .YP-; U12 = S12(S,zK*KS12 + I-)-12, S12 is a semi-unitary operator carrying .YW'- into .Ye+. To solve the problem it suffices to vary the operators K, S12 and S21, and also the number of examples of the space .W, and to put V = w+ (V ). S21
8
Let (m, n) be an arbitrary pair of non-negative integers. Give an example of an unbounded closed (J1, J2)-isometric operator V with
=.;f and m=dim (/+(v), n=dim .Jt4-(/)
6;v
(essentiality of the condition for boundedness of V in Exercise 6). Hint: Use the device given in the hint on Exercise 7. 9
Let V be a (J1, J2)-isometric relation, V1, = 1 v, .;?v = . 'v, and let in, n, p, q, r be
the same as in Exercise 6, and s = dim Ker V -'. Then p + q = in + n + r + s (Azizov, E. lokhvidov and I. lokhvidov [ 1] ). 10
Let .YP be a J-space, T a linear relation, and nr(T) = dim .-IT- rt. Prove that if T is a J-Hermitian linear relation, 1 ;e 3 , and .-1T- tt and WT- r T are subspaces in.YI , then
nl-jT) = nr(JT) + nl-(JT) + dim Ker(T- I) + dim Ker(T- ('I). In particular, if is a point of regular type for T, then the condition that 'WT- r t be closed may be omitted (Azizov, E. Iokhvidov, I. Iokhvidov [1]). Hint: Use the results of Exercises 6 and 9 and also Theorem 1.14. 11
Under the conditions of Exercise 10 let T be a J-Hermitian operator. Prove that T = TT if and only if
n1(T) + nt(T)=dim Ker(T- J)+dim Ker(T- ('I) (Azizov, E. Iokhvodov, I. Iokhvidov [1]). Hint: Use the results of Exercise 10. 12
Prove that
if
T is
a
J-non-expansive operator
and I O ap(T),
then
K;(K;'(T)J)=w+(T).
Hint: Use Exercise 33 on 2.§4 and Corollary 1.15. 13
Let 1=(o[+]12 and 11=('J1)o[+] (1'1) [+] ('J1)2 be the 1 and subspaces
decompositions of the .H = 1'o [+] 1i [+] 12 O+ (11)o [+]
(.YP+,.YP-)-
'j +
let
[+]('/'1)2, and let J= 11 Ullbi=1 be the matrix representation of J (see Exercise 5 on 1.§10). Verify that a bounded
252
5 Theory of Extensions of Isometric and Symmetric Operators operator X with 1'x = .0 represented by the matrix X = 11 X,jII 6;= I will be a J-bi-extension of the operator Iv = I I C1 if and only if 11 X;ijI i;= I = I,, X;j = 0 when i=4,5,6, j=1,2,3; X2;=0 when j=4,5,6; .3 X6jC Rj6b and X31= - J36 FF6' X6i, j = 4,5; 'I6 - x66 Cj66 and X36 = J36 46V - 66);
X44 = I, and all the remaining X;; are arbitrary (Azizov [14] ).
Hint: Write out the matrix X` and use the inclusion If C X' derived from the inclusion in Proposition 1.12. 14
Let .' be a J-space, 9 a subspace of ., and J = II JJlMM ?j= 1 be the matrix representation of an operator J relative to the decomposition .r° = V Q+ V -L. Prove that an operator Y( E :i) will be a J-bi-extension of the operator I, if and only if there is operator an Zzz : V2 - 1-1 such that 00ap(I2 + J22 +(I2- J22)Z22) and Y= II
11
Y1i
%.i= 1, where YI1 = Iv, Y21 = 0,
Y12= -2J12(I2-Z22)(12+ Jzz+(h - J22 V22)
Y22=(I2- Jzz+(II+ Jzz)Zzz)(12+ J22 +(I2- Jzz)Zzz)-' (Azizov [14]). Hint: Verify that w+ (Li) = Ls; use Proposition 1.11 and establish that biextensions of the isometric operator If coincide with the set of operators representable by matrices Z= jI Z;ijI ij= 1, where Z1l = Iv, Z12=0, Z21=0, and Zzz : 9)-L - Y 1; verify that the condition Z E J is equivalent to the condition 00ap(I2+ J22 +(12- J22)Z22) and that Y=w+(Z). 15
Prove that under the conditions of Exercise 14 the operator Y will j-bi-nonexpansive (respectively, J-semi-unitary J-bi-non-expansive, J-unitary) if and only if Z22 is a contractive (respectively, semi-unitary, unitary) operator with 1'z22 = 9 and 0 E p(I2 + J22 + (12 - J22)Z22) (Azizov [14] ). Hint: Use the hint on Exercise 14 and Corollary 1.7.
16
Let V be a (J1, J2)-isometric operator which has a continuous (JI, J2)-biextension Vo with 9 vo = . , a projectionally complete range of values, and Ker Vo = 161. Prove that the operator P with VV= W 1 is a (JI, J2)-bi-extension of
the operator V if and only if V=I.,,Vo(VCVo)-', where I.,r is a certain J2-biextension of the operator I.,? = 12 v6; V'i.,,. = y f2 (Azizov [ 14] ). Hint: Use the fact that Vo Vo is a linear homeomorphism of the space .Yi and is a J1-bi-extension of the operator I I Vv. 17
§2
Prove that if, under the conditions of Exercise 16, the operator Vo is (JI, J2)-semiunitary, then the formula given there can be rewritten in the form V = I.,QVo, and if Vo is a (JI, Jz)-unitary operator, then V is a (J1, J2)-unitary (respectively, (JI, Jz)-semi-unitary, (JI, J2)-bi-non-expansive, (JI, Jz)-bi-non-contractive) operator if and only if i.w belongs to the corresponding class (Azizov [14]).
Extensions of standard isometric and symmetric operators
; _ .YP; Q+ W1 be Ji-spaces (i = 1, 2). In this section we present 1 Let another approach to the construction of a theory of extensions of (JI, Jz)isometric and J-symmetric operators and we describe all their (J1, J2)-biextensions and J-bi-extentions respectively.
§2 Extensions of standard isometric and symmetric operators Definition 2.1: We shall call an operator V: .W'1
253
W2 a standard operator
and write V E St(."1, ."2) if V is a closed continuous and continuously invertible (J1, J2)-isometric operator. In particular, if ._W, = .W'2 = ,Y is a J-space, we shall write V E St X. We point out that the condition VE St(, '1, 72) implies that Yv and Rv are closed. Using Zorn's lemma it is easy to prove that every (J1, J2)-isometric operator
V admits extension into a maximal (J1, J2)-isometric operator V, i.e., V now has no non-trivial J1, J2)-isometric extensions. Less obvious is the assertion that if V E St(7,, M2), then among its maximal extensions there is a V E St(7?'i, M2). Theorem 2.2 is devoted to the investigation of this question; in the proof of this theorem a construction for the corresponding extension is given which will be used more than once later.
Theorem 2.2: Let VE St(1,.2). Then it has a maximal (J1, J2)-isometric extension VE St(,W1, '2).
First of all we observe that if 9I-1 and RI1l are regular subspaces, then, as is easily verified, the operator V has an extension V'E St(,,, .Y2) such that its deficiency subspaces 9 vl] and ,3W are regular and either at least one of them is (B) or they are uniformly definite and have different signs. The maximality of such (J1, J2)-isometric operators is obvious. Consequently the theorem will be proved if we verify that every operator VE St(,,,M2) admits extension into an operator V' E St(.7,, W2) with regular deficiency subspaces VVJ and We consider the (,7i , W1 )-decomposition of the subspace 9v (see 1.Definition 10.3): RV).
V V = C/o [-H 91 [+] ci2
(2.1)
and the (.7?', .z )-decomposition of the subspace Rv:
Rv=Ro[+] W, [+]12.
(2.2)
Without loss of generality we shall suppose that (2.1) and (2.2) are connected
by the relations VCA, = R,, i = 0, 1, 2. Indeed, Ro = V90, since a (J1, J2)isometric operator maps the isotropic part go of the subspace Vv on to the
isotropic part Ro of the subspace Rv. But if even one of the relations R, = V '1 (i = 1, 2) is infringed, we proceed as follows. Taking 2.Corollary into account we shall suppose, without loss of generality, that v((' 1 n .-i) C .Y( and we consider the decomposition
4.13
Jv= Vo [+] V-'R1 [+] V-'R2.
(2.3)
Hence 2= [4-J ]J , (1 C V-'R1, V-1R1 = c1 [+] V2, and where 92"= V-'R2 By virtue of 1.Theorem 10.4 we can find a maximal 11
5 Theory of Extensions of Isometric and Symmetric Operators
254
uniformly definite dual pair (Y1', Yi) such that (2.3) will be the (2l+, Y; )-decomposition of the subspace 9v. It only remains to put Let
Q'V =(V1)0[+]W)I[+]W)2 and
YV = (
1)0 [+] (
1)1 [+] (R1)2
be the (771 , .W )-decomposition and the (.,Yz , .,Yz )-decomposition of subspaces V 'V and i.Rt respectively. Then relative to the decomposition
the
.Iy1 = /0 [+] t1 [+] 92 O (V')o [+]W)1 [+](d 1)2 the (respectively, '2 = ,!0 [+]4?1 [+] 2 O 0?1)0 [+] 0?l)1 [+] (.1)2) operator J1 (respectively, J2) can be expressed in accordance with Exercise 5 on 1.§10 in the form of a matrix J1 = I Jtil) I61 (respectively, J2 = 11 JJ(jzt1I -=1.
The components of the expansions of the matrices J1 and J2 have the properties indicated in Exercise 5 on 1.§10. Thus, for example, J16) _ 3>2)1i2V36),
(131) - A3>2)"2 06), and j162) = (I 2) - J
where V}6) is an isometry
mapping (d 1)2 on to d2, and V36) is an isometry mapping (1)2 on to JV2. We introduce the notation V;+1 = V I d; (i = 0, 1, 2) and define an extension V' of the operator V in the following way
V'Idv=V; V' X1)0
V4:(V1)0- (1)0,
V4= Jgl)V
1J14);
2 0 W)2, V6X= V36X+ V6X, XE ((1)2,
V' I(d1)2 = V6: (d1)2 V36X E 32, V6X E (3q1)2,
V36 = [ (132) + (J2) -
J33)2)1/2)- I J33
V*- I
(Jill - J33)2)-1)J33)] 116), V6 = V36)* V3 *-1 V36).
The operator V' is continuous and continuously invertible; (I V' = V O [+] (I1 [+] V2 @ (V 1)0 [+] (V' )2, and
RV'=,3? [+]31[+]+J?2(@ ( 1)0[+](91)2.
Moreover, V'(do 0 (V1)0) = ,Ro O ( 1)0, V' d1 = ill, and V'((I2 O (d1)2) = W2 +O (-?')2. In accordance with Exercise 4 on 1.§10 the subspaces and `do O+ (d 1)o, d 1, and 92 O+ (V')2 (respectively, 4?o O+ ( 1)0, 1,
.R2 O+ (1)2) are pairwise JI-orthogonal (respectively, J2-orthogonal), and these subspaces are invariant relative to J1 (respectively J2). Therefore V' will be a (J1, J2)-isometric operator if and only if the operators V' `/b O+ (V 1)0, V' (I 1, and V' I V2 O+ (I 1)2 are (J1, J2)-isometric. Since V' V I = V I (I1, the operator V' (l is (J1, J2)-isometric (by hypothesis). Let XI E ('o, I
§2 Extensions of standard isometric and symmetric operators
255
X4 E (`I' 1)0. Then V' (x1 + X4) = V1 X1 + J4f) V,'-' Ji4')x4, and
J2V'(x1+x4)= J13VIx1+ Vl* 1Ji4X4. Consequently 1
[V'(X1 + X4), V'(XI + X4)12 = (Jf4)*Xl, x4)1 + (J14)x4, x1)1 = [xl + X4, X1 + X41111,
i.e., V' I 1o ®+ ('1)o is a (J1, J2)-isometric operator. We verify that the operator V' I V'2 O W)2 is also (J1, J2)-isometric. To do this we shall suppose, without loss of generality, that W1 = V2 O+ (91)2, %P2 = R O+ (,W1)2, and we shall prove that V' is a (J1, J2)-unitary operator, i.e., V'* J2 V' = J1. It can be verified immediately that the operator T = V' * J2 V' has the matrix representation
I
T33
T36
T63
T66
where T33 = V3 Jf3' V3,
T63 = T36 = V3 J33 V36 + V3(I32) - J33
2)1/2 I/6,
and T66 = V36 J33) (V36 + V6 * V36 *(I32) - J33)2)1/2 v3(2 + V36 (I32)
6
J(3)2)1/2 "36
v,6 -
V6 *
V3(6)* j3(2) 3
32
V2
V6.
Since the operator V3 is (JI, J2) isometric, we obtain T33 = J431). We substitute the values of the operators V36 and V6 in the expressions for T36 and T66, and after some elementary transformations we find that T36 = Jf b) and T66 = 46), i.e., T= J1. Thus V' I 942 O+ (ft')2 is a (Ji, J2) isometric operator and
V' E St(.01,.W'2). Moreover, (y1)1 and (.R1)I are its deficiency (regular)
subspaces and therefore V', and with it also V, has among its maximal (J1, J2)-isometric extensions also operators VE St(.'1,.-W2).
For an operator VE St(.Iy1,.2) to be a maximal (J1, J2)isometric operator it is necessary and sufficient that at least one of the Corollary 2.3:
following conditions shall hold: a) VVl] = )B), i.e., V is a (Jr, J2)-semi-unitary operator;
b) .1111 = )B), i.e., V-' is a (JI, J2)-semi-unitary operator, c) VI1J and R11] are uniformaly definite subspaces of different signs.
The sufficiency of each of the conditions a)-c) for maximality of a (J1, J2)-isometric operator V is obvious. Now let V E St(.1P1i .2) be a maximal (J1, J2)-isometric operator. From the course of the proof of Theorem 2.2 it follows that #l and ow;,-'l are regular
256
5 Theory of Extensions of Isometric and Symmetric Operators
subspaces. Let us suppose that none of the conditions a)-c) is satisfied. Then there are B ? xo E and 9 * yo E t such that [xo, xo] = [yo, yol. We define an operator V' : Lin (V v, xo) - Lin) i v, yo) , V C V', V' xo = yo. It is easy to see that V' E St(.°1i '2)-and we have obtained a contradiction. J
Corollary 2.4: Every operator V E St(. 1, .W'2) admits a (JI, J2)-bi-extension Vo such that :91 vo is a projectionally complete subspace and at least one of the operators Vo or Vo has a trivial kernel.
By virtue of Theorem 2.2 we can suppose that VE St(. '1i'2) is a maximal (JI, J2)-isometric operator. We now use Corollary 2.3 and Proposition 1.11. If V satisfies condition a), then we put Vo = V; if V satisfies
condition b), then we put Vo = (V -' )`; if V satisfies condition c) and dim qV] < dim 5t1then as Vo we take any extension of the operator V on to WI which maps
I, then as Vo we take any extension of the operator Von
to WI which mapsY!A'I injectively on to a subspace in RV'I, and if dim IL] > dim 9l `l] , then we put Vo = (V-' )`, where V-' is an arbitrary extension of the operator V subspace in ILl
on to W2 which maps 9 I'l injectively on to a
Thus, for the description of all (JI, J2)-bi-extensions of an operator VE St(.1,.2) the results of Exercises 13-17 on §1 can be used.
2 We pass on now to study the problem of the possibility of (JI, J2) unitary extensions of operators VE St(. 1, JY2). Example 2.8 given below of an operator V E St(.W1, ."2) with 9v = Rv (and therefore 9 fr1 = 3 *) shows that, in contrast to the Hilbert space case (see, e.g., [I] ), in the case of a Pontryagin space (see Exercise 3 below) information only about the properties of 9 Vll and f'l is not always sufficient to make conclusions about the possibility of extending an operator V into a (J1, J2)-unitary operator. But first we introduce some preliminary materials. Definition 2.5: Let (91+, 2-) be a maximal uniformly definite dual pair in a J-space . , and suppose the subspaces A"+ E -ff+ and ./I-- =.4'[+-,I (E 11-)
contain no infinite-dimensional uniformly definite subspaces. Then the number v (.,l "+) = dim(./l "+ n Y+) - dim(,il n Y-) is called the index of the subspace . 4'+.
Lemma 2.6: In Definition 2.5 v (. 'l '+) does not depend on the choice of the (Y+, Y-). maximal uniformly definite dual pair Let
P1( l"+)=dim(.,l"+ fl.,Y+)-dim(./l"_ n.w- )
§2 Extensions of standard isometric and symmetric operators
257
and
v2(. '+) =
dim(.'{"+ n q+)
- dim(.'1i. fl g'-).
We verify that v, (,/1'+ ) = P2(,4'+ ).
Let K+ be the angular operators of the subspaces A-+ and Q± be the angular operators of the subspaces 9?± . From Exercise 17 on 1.§8 and Exercise 13 on 2.§2 it follows that K+ is a 4)-operator and I K+ I = I+ + S, where S E Y.. From Exercise 14 on 2. §2 it therefore follows that K+ - Q+ is also a 4)-operator and ind K+ = ind(K+ - Q+). It remains to observe that ,4'+ n ± = Ker K+, A"± fl 2± = Ker(K+ -Q±), and K+=K*, K+-Q+=K*-Q*, i.e., -ind K+ and v2 (-4"+) = - ind(K+ - Q+ ).
Lemma 2.7: If VE St(.Y(1,.Y2), d'vE M+(JY1), .9lvE,11+(."2), and 1'v, Rv, gW and 91VI contain no infinite-dimensional uniformly definite subspaces, then V admits a (J1, J2)-unitary extension if and only if v(Vv) = v (91 v ). Moreover every maximal extension V' E St(.Y1, '2) of the operator V will be (J1, J2)-unitary. Let 17 be a (J1, J2)-unitary extension of the operator V. We introduce in
.Y'2 a scalar product (x, y)2' = (V-'x, V-'y)i (x, yE.)2) equivalent to the original one. Since [x, y]2 = (Jzx, y)2, where Ji = 17J1V-' = Jr', the components of the new canonical decomposition 2 = 'Y2" O+ .VZ 'will be the subspaces ,Yf F.01'. Consequently dim(vv n i ) = dim(V(cv n .)Yi ) = dim(3 v n ,y2,,) and dim(91*1 fl e-) = dim(V(Qo ] fl Yi 1)) _ dim( V, fl W '), and therefore v(V v) = v(,R v). Now let P(Vv) = v(Rv) Without loss of generality we shall suppose (see the proof of Theorem 2.2) that the decompositions (2.1) and (2.2) of the subspaces cv and Rv are connected by the relations VCA; = Ri, i = 0, 1, 2.
Since c1 = Iv n .iYi and 1 = 91v fl .w2+, we have dim(iv n i) = dim(3ly n . w2+). Consequently dim(cv, n W,-) = dim(f1w fl. l-) = -v(/v)+dim(Vvn -wi v(JRv)+dim(.91vn,Yzdim(.RVI n.YZ ) = dim(MV n . i ).
Following the scheme of the proof of Theorem 2.2 we deduce that the operator V admits an extension V' E St( 1, . 2) with 9W n .01- and wv n Wi . Hence V VJ and [1I are negative subspaces of equal finite dimension and therefore it is now easy to construct a (J1, J2)-unitary extension of the operator V' and it will also be an extension of V. Let V' E St(.,*P1, -02) be an arbitrary maximal (J1, J2)-isometric extension
of the operator V, and suppose that V' is not a (J1, J2)-unitary operator. Since V;,1l and ,j-] have the same sign, we conclude from Corollary 2.3 that we can, without loss of generality, suppose the operator V' to be (J1, J2)-semiunitary and dim R Vi] ;4 0.
We consider the J2-space ,Y2= V'. 1. From what has been proved, P(Vv) = v'('), where P'(') is the index of Mv in .W''. But by hypothesis
258
5 Theory of Extensions of Isometric and Symmetric Operators
and it is easy to see that v(3'v) = P' (3 ) - dim .91(1j-so we v(V v) = have obtained a contradiction. We pass on now to the construction of an operator VE St Y' with (v= ,3v, but which does not have J-unitary extensions.
Example 2.8: Let V1, LA2, (!Yl)2 be infinite-dimensional separable Hilbert spaces. We form the space ,Y = 91 Q+ CA2 Q+ (GO ')2 and introduce in it a J-metric (see 1.Example 3.9) by means of the operator
J=
J22
0
0
J33
0
V36 (I3 - J33)
0
(J3 - J33 1/2
)1/2
V36
,
- V*J03 V36
constructed relative to this decomposition of the space ye, where J22 = 12, J33 E Y., 0 < J33 < 13, and V36 is an isometry mapping ('W1)2 on to 92. Let be the eigenvalues of the operator J33, and let X1 >, X2 >, limk-., Xk+1/Xk = a > 0; let [ fk) be an orthonormalized basis in 92 composed of the eigenvectors of the operator J33: J33fk = Xkfk (k = 1, 2, . . .); and let [ ek 11 be an orthonormalized basis in 91. We define on 9 v = 91 O+ 92 an operator V E St W by putting V e k = ek+1 (k = 1 , 2, .
.
.),
Vf1= Xl el,
V f k = Xk+1/Xkfk
(k = 1, 2, ...).
By construction My = cv. The operator V admits no J-unitary extensions V, for if it did the operator VI V2 (E St(,'1, ,1W2)) acting from the J1-space .)r1 = C42 O (4 i )2 into the J2-space W2 = Lin [ el) + CA2 O+ W)2 with Ji = J I Yei, i = 1, 2, would have (J1, J2)-unitary extensions, and this is impossible by virtue of Lemma 2.7 since P(92)= 0 and v(Vi2) = 1.
The above example shows that in answering the question whether an operator VE St(.'1,,W2) has (J1, J2)-unitary extensions one has to take into account not only the characteristics of the spaces CA I1l and RI'l but also the `action' of the operator V. Let be a maximal uniformly definite dual pair in Ye1, let V V = C/o [+] 1/ 1 [+] V2 be the ('l', 21 )-decomposition of the domain of definition of an operator V E St(,*'1 i 46), and 91 v = 3o [+] V!Y 1 [+] Vc12 be the decomposition of its range of values. We choose in .-W2 a maximal uniformly
definite dual pair (Y2, Yz ) such that Vw v fl 99i ) C Y27+, and we introduce the following constants v+ (V) and v _ (V ): if there is in V[,1l or in an infinite-dimensional uniformly negative (respectively, uniformly positive) subspace; (2.4) dim(V I-L] fl Y ) + dim( V/2 fl £Z) - dim(. 1[,1] fl Pz ) otherwise. 0
v±(V) =
§2 Extensions of standard isometric and symmetric operators
259
Remark 2.9: If there are no infinite-dimensional uniformly definite subspaces in Vv, v, 9111 and R(,1], and if 9v E,.tl+(.W'1), and i?vE tl+(,°z), then
v_(V)=0 and v+ (V) = v(Av)- v(C'v). Our immediate purpose will be to prove that v±( V) are independent of the choice of the dual pairs (-Ti', 2 ), i = 1,2. We preface this with the following proposition.
Lemma 2.10 Let . be a J-space, f a subspace of it, and let U be a J-semi-unitary J-bi-non-contractive (respectively, J-bi-non-expansive) extension ;e (0) implies that there is in c(1) of the operator I I V. Then the condition an infinite-dimensional uniformly negative (respectively, uniformly positive) subspace.
Let !J = go [+] c1 [+] 92 be the )-decomposition of the subspace J, and let U be a J-semi-unitary J-bi-non-contractive operator. Since 90 and Col] are invariant relative to the operators U and U`, it follows in accordance
with Exercise 20 on 2.§4 that the operator U induces in the i-space _ Col]
!I o
a
J-bi-non-contractive
Jsemi-unitary operator
U1 9 = J1 9, where I = 9/ go. Moreover, ML'I ;;d
(0)
U,
and
if and only if
U ;e (0); and 9[1] contains an infinite-dimensional uniformly negative subspace if and only if (1] _ 9 [1]' go has this property. Therefore we shall suppose without loss of generality that go = (0), i.e., ', and with it also iI [1] and g', are non-degenerate subspaces. In accordance with 2.Proposition 4.14 we have for the J-bi-non-contractive operator U that M I l (= Ker U`) is a uniformly negative subspace, and so without loss of generality we shall suppose that 91Il C -W-. We shall assume
that W # (0) and that, contrary to the proposition, V[l] does not contain infinite-dimensional uniformly negative subspaces, i.e., (see Exercise 7 on 1.§6) the Gram operator of every non-positive subspace from c l1] is completely continuous. Let U = II U jII ? j=1, J = II J;i ?i= 1 be the matrix representations
of the operators U and J relative to the decomposition W = SI (1 91 . Since mil[ 1 C Y and ci CRC, we have J& C V`(= Jc [1] ), and U I = 9lu2, n ci 1 = i i I n c 1. By direct calculations taking Exercise 7
on 1.§3 into account we verify that the fact of the operator U being J-semi-unitary is equivalent to the relations a) .V1, - u22 C . IrZZ, U12 = - J21 Jzz' (Iz - U22); b) Uzz Jzz' Uzz Jzz = Iz.
Since Ker Uzz =
u,z n `I' 1, and condition b) is the condition for the operator be to Jzz-semi-unitary, so
( Uzz) ` = Jzz' Uzz Jzz
(Ker U2*2)111). Let CA1 = c+ [+] V_' be the canonical decomposition of 'J 1, and let P+' be the projectors on to V +, P+ + P_ = 12. It follows from the relation a) and the complete continuity of J2'21 Y_ that the operator
T= P_Jzz'(I2- U22 )J221 ci_' =I_ - P_Jzz'UzzJzzI f'
260
5 Theory of Extensions of Isometric and Symmetric Operators
is completely continuous, and therefore P_ Jzz' U22 J22 9' = I_ - T. Since Jzz' U22 J22 is a J22-semi-unitary operator, we have Ker(P_ Jzz' U22 J221 CA') _ (0) and therefore i?j - r= LA-. Hence we conclude that A2' U22 J22 C4' is the
maximal non-negative subspace in 9 1. But the subspace Ker Uzz (6 {6) ), which is J-orthogonal to ,?(uiz)', is also negative-and we have obtained a contradiction. The case when U is a bi-non-expansive operator is proved similarly. Lemma 2.11:
In Definition (2.4) v±(V) do not depend on the choice of the
dual pairs (?; , 2' ), i = 1, 2. We shall verify that v+( V) does not depend on the choice of the dual pairs the argument for v- (V) is similar. If there is an infinite-
dimensional uniformly negative subspace in mfr], then v+(V)=0 by definition does not depend on the choice of the dual pairs. Suppose all the uniformly negative subspaces in Vl] are finite-dimensional. Without loss of generality we shall assume that the decompositions (2.1) and
(2.2) are connected by the relations Vii; = 3;, i = 0, 1, 2, and therefore the
constant p+(V)v, calculated with respect to the dual pairs i = 1, 2, coincides with the difference dim(9f1' fl;)-dim(3[1I fl .z). We bring into consideration a J,-space ;, and a J2-space 46, putting , y , [ O+ ].
1,011
0]0 1,0-
,W2[O] I
where dim
-_I
P, I,
j, IYe
G)
_
±K± if v, <0,
i2I,'Y±
[OJ
+, Ye2[O+],W+,
and ,Y+
if v, > 0,
J, .W+ = I+
+,
Jzl ye+I+ is
=±1±
if v,>0, if v, 50,
an infinite-dimensional space with
dim + > max(dim ,, dim .)Y21. It is clear that the constant v+(V) = v, (see (2.4)) constructed relative to 4e ,t, . e, ), i = 1, 2, is equal to zero, and it is
sufficient for us to verify that it does not change for any maximal uniformly definite dual pairs (S,+, i = 1, 2 containing (.YP+, .Yfo ), where .moo = if Y- enters into the considered space ; , (i = 1, 2), and No = (0) otherwise. Since J, = 0, we can show, by repeating the arguments used in the proof of Theorem 2.2, that the operator V has a (J,,12 )-unitary extension V,. Let JZ be 7), the constant (2.4) calculated relative to some other of the dual pairs i = 1, 2, indicated above. Again reverting to the proof of Theorem 2.2 and
using Lemma 2.7 it can be proved that the operator V has a maximal (J,, j2)-isometric extension V2 E St(.;,,.ie2), and that one of the operators V2 or VZ ' is defined on the whole space, and the orthogonal complement to its range of values is negative and has dimension I v2 1. Let V2 be such an operator (for Vi- ' the argument is similar). Then the operator V,- 'V2 is a j,-semi-unitary J,-bi-non-contractive extension of the operator I I i v and 9 [ J contains no infinite-dimensional uniformly negative subspaces. By virtue of Lemma 2.10
§2 Extensions of standard isometric and symmetric operators
261
V,-'V2 is a Jl-unitary operator, and therefore V, is a (J,, j2)-unitary operator, which implies the equality j72 = 0.
We now pass on to the main result of this section. Theorem 2.12:
An operator VE St(.°,, 42) admits a (J,, J2)-unitary exten-
sion V if and only if v+ (V) = v _ (V) = 0 and there is an operator V, E St(.',, J '2) which maps !2V1' on to &V1I
Let V be a (J,, J2)-unitary extension of the operator V. We then put V 9 V = V, E St(°,, . 2) and it is clear that V, 9 Vl' _ M Vil. The equality v+ (V) = v_ (V) = 0 follows from Lemma 2.11 since this equality holds for the
dual pairs (.01', ,i) and (V1+ , V.Wj ). The converse assertion will be proved in several stages. a) Suppose first that v+ ( V) = v _ (V) = 0 and that 9 VI and V1' contain no infinite-dimensional uniformly definite subspaces. Repeating the arguments
used in the proof of Theorem 2.2 we obtain that the operator V admits (J,, J2)-unitary extensions. b) Let V1' contain infinite-dimensional uniformly definite subspaces of both signs, let V, E St(.W,, 02), V,9V1' = M V ], and let V'(E St(JY,,.'2) be a maximal (J,, J2)-isometric extension of the operator V. We go into the case when V' is a (J,, J2)-semi-unitary (J,, J2)-bi-non-contractive operator and we show that it can be `touched up' into a (J,, J2 )-unitary extension V of the
operator V. The remaining cases are verified (taking Corollary 2.3 into account) by the successive application of a similar procedure. Since 9 Vi] contains infinite-dimensional uniformly negative subspaces, MI'I (= V, 9 Vl' ) also has this same property. Let q2 be such a subspace 2', = Vi '22. Since let from 9LVlI R[1] and containing
and 9?, and V' 2'1 [+] M V,I are uniformly negative subspaces, there is an operator V(E St(.Y,, '2) which maps 2', on to V' 2', [ f ] M VI. It only remains to put dim 9, = dim( V' 2', [+] 91 V1 ),
(V'x,
Vx= (V, x,
xE 2'1'] X E 9?,
c) Now suppose that V V I contains infinite-dimensional uniformly definite subspaces of one sign and only finitely-dimensional ones of the other sign, and v+ (V) = v_(V) = 0. In this case a combination of the arguments of stages a) and b) is applied. Remark 2.13: In stage a) in the proof of the converse assertion in Theorem 2.12 the existence of the operator V, was not used, and in stages b) and c) in
this proof the operator V, was used only in the search in ( VI'' for an infinite-dimensional uniformly definite subspace of the same dimension as the corresponding subspace chosen in J?VrI
262
5 Theory of Extensions of Isometric and Symmetric Operators
If under the conditions of Theorem 2.12 91V1] = Vo [+] CA [11 where '/b = vv n V v`], and t [1[ ' is a projectionally complete subspace in .YP1, then the existence of a (J,, J2)-unitary extension V is equivalent to the In equality In Corollary 2.14:
El It follows from the conditions on 91 ] that 9 v = Vo [+] 9', where V' is a projectionally complete subspace (see Exercise 7 on 1.§7), and therefore ?v and have property. a (_ .Ro similar Let v1[
(`e[1[)+ [4](t'f'-))_ and W[-Ll' = (,W[-LI)+ [+]By hypothesis , and (,[1[)+, (.i?[1) + are
dim V'o = dim 3 o, dim(V[1[)± = dim(M [1) +
uniformly definite subspaces, and so there is an operator V, E St(.,, .02) mapping VI'] on to V Taking as corresponding maximal dual pairs q ), i = 1, 2 the pairs of spaces 'i 3 (V [1) ± , 2'z 3 (.i [1]) ± we obtain that v+ (V) = v _ (V) = 0. It only remains to apply Theorem 2.12. The converse assertion is obvious.
3 We introduce the concept of x-regular extensions of operators VE St(. ,,.-N'2), and we give a criterion for their existence. But first we prove the following simple proposition.
Suppose that P is a (J,, J2)-non-expansive extension or a (J,, J2)-noncontractive extension of an operator V E St(.W,, W2) on to the whole of .01, that V admits a (J,, J2)-polar decomposition V = WR, where the operator W is a (J,, J2)-semi-unitary operator or is conjugate to a (J,, J2)-semi-unitary operator, and the J,-selfadjoint operator R satisfies the conditions R2 = V`V and a(R) C [0, co). Then W is an extension of the operator V Let x E V v. Then R 2x = 17']7x = x, and since - 1 E p (R) we also have Rx = x. Consequently W 3 V.
2.15
Remark 2.16: If under the conditions of Proposition 2.15 the operator W is (J,, J2)-semi-unitary, then it is a (J,, J2)-bi-extension of an operator V (see
Proposition 1.10). We note also that if V is a J,, J2)-bi-non-expansive operator, then by virtue of 4.Theorem 1.12 it admits the decomposition indicated in Proposition 2.15, and moreover W is a (J,, Jz)-bi-non-expansive operator and therefore (see Proposition 1.10) W is a (J,, J2)-bi-extension of the operator V.
Definition 2.17: We shall say that an operator V E St(.",, .W2) admits x-regular extensions (and when x = 0, regular extensions) if there is a Pontryagin 7r,-space II, with x (< oo) negative squares such that the operator V admits (J,, J2)-unitary extensions V: 4i @+ fI - .)Y2 0+ II,,,
(here 7r = J in the space II).
J1 = J1 Q+ 7rx,
J2 = J2 (+ 7r
§2 Extensions of standard isometric and symmetric operators
263
Theorem 2.18: Let v fl St(.r,,.2 ). Then there is a J-space.r such that the operator V admits a (J,, J2)-unitary extension
V: .r, O+ Jr ' r2 (@ . i,
Jr = J; (D J,
i = 1, 2.
Moreover the following conditions are equivalent: the operator V admits a) regular extensions; b) x-regular extensions; c) (J,, J2)-bi-non-expansive extensions.
We shall suppose that VE St(.0,,.r2) is a maximal (J,, J2)-isometric operator. From Corollary 2.3 its deficiency subspaces and W-'] are projectionally complete. To prove the first assertion it is sufficient to choose as Jr the J-space .-W® D Jr - with infinite-dimensional components Jr
dim Jr ± > max (dim 001'1, dim 11 V] }, and to use 2.14.
We pass on to the proof of the equivalence of the conditions a), b), c).
a) - b) This is trivial.
b)-c) Let V:.r, O+ fl -.r2 H, be a x-regular extension of the operator V, let IIx = TI+ O+ II-, dim II_ = x < co, and let P; be the J;-orthoprojectors from i on to the J space .r, = *, O+ II_, J; = J; O+ - I_, i = 1, 2. It follows from Exercise 34 on 2.§4 that P2VPI I.h1j is a (Fi, J2)-bi-nonexpansive operator, and since Vv C i, R v C .Vz, it is an extension of the operator V. We make use of Remark 2.16 and suppose that, for example, the operator V- 1 has a (J2, Ji )-semi-unitary (J2, JI)-bi-non-expansive extension V' (if V has a Ji, J2)-semi-unitary (JI, J2)-bi-non-expansive extension, then
the argument is similar). Let V' = Vij' - --, be its matrix representation relative to the decomposition .Z = .'2 U H_, Jr( = . , O+ U _ , and let x+ = dim MV1'. Then the operator V' = I Ker Vz, is a (J,, J2)-isometric exten-
sion of the operator V -', and since dim Jz Ker V21 < co, so V v and v, decompose into the direct sum of isotropic and projectionally complete subspaces. Let In 1 (xo, x+, x-' ). Then In
_(xo,x_+x+,x_
Since II _ C V ,] and II _ C. of ,] , we have In(
[,'flr2)=(xo,x+, x_ -x)
and
-x), and moreover
dim(co[,1'fl.W2)=xo+x++x_ -x < 00
.
Now, carrying out an argument similar to that used in the proof of Corollary 2.14 we obtain that V -' has a (J2, J, )-semi-unitary (J,, J,)-bi-non-expansive extension, the (J2, J, )-conjugate to which will indeed by the required extension of the operator V.
264
5 Theory of Extensions of Isometric and Symmetric Operators
c) - a) We again make use of Remark 2.16 and we shall suppose that, for
example, the operator V-' has a (J2, Ji)-bi-non-expansive (J2, Jl)-semiunitary extension W. Since (see 2. Remark 4.29 and 2. Proposition 4.14) 5$wl is a uniformly positive subspace, so by virtue of Corollary 2.14 the operator W has a (J1, J2)-unitary extension W : ,7 '2 O+ . Je1 O+ .', J; = J1 O+ I, i = 1.2, where , is an infinite-dimensional space with dim .c > dim 91wl. Con-
sequently W-1 is a regular extension of the operator V.
In conclusion we introduce some definitions concerning J-Hermitian operators, and we formulate in Exercises 6-13 the corresponding results 4
obtained by applying as above the Cayley-Neyman transformation. Definition 2.19: We call an operator B a J-bi-extension of a J-Hermitian operator A if A C B and the graph I'A is isotropic in I'B relative to the form
[(x1, x2), (Y1,Y2)]r= i([x1,Y21 - [x2,Y1] (2.Formula (1.4)).
The set of closed J-Hermitian operators A with points of regular type (Xo, Xo) (Xo 5;d ko) will be called Xo-standard and will be denoted by the symbol St(.h"; Xo) (= St(Jf; ko)).
Definition 2.20: We shall say that an operator A E St(,'; 4) admits a x-regular (0 < x < cc) Xo-standard extension if there is a Pontryagin ax-space II with x negative squares such that in the J-space = W C+ II
(J = J @ H.) a J-Hermitian operator A has a J-self-adjoint extension A E St(,*; Xo). When x = 0 such an extension will be called a regular extension.
Exercises and problems 1
Let V E St(.r#1, k2). Prove that v, (V-1) = - v, (V) (Azizov). Hint: Use Formula (2.4) and Lemma 2.11.
2
Suppose that VE St(./P1, IY2), that V has at least one (J1, J2)-unitary extension, and that ciH and RI-LI contain no infinite-dimensional uniformly definite subspaces. Prove that then every maximal (J1, J2)-isometric extension V' (E St(. 1, .,Y2 )) of the operator V is (J1, J2 )-unitary (Azizov). Hint: Prove that x ± (V') = 0 for any extension VE St(.10'1,./<'2) of the operator V.
3
Prove that V admits a (J1, J2)-unitary extension if and only Let V if XI = x2 and dim V47L] = dim . H (cf. [XIV]). Hint: Use I. Theorem 9.11 and Corollary 2.14.
4
Prove that under the conditions of Corollary 2.14 the projectional completeness of
v'lll' is essential. Hint: Use Example 2.8. 5
Let A be a J-Hermitian operator, B an extension of it, and VB = ./l'. Prove that B will be a J-bi-extension of the operator A if and only if A C B. Hint: Use 2. Proposition 1.3.
§2 Extensions of standard isometric and symmetric operators 6
265
Let A be a J-Hermitian operator, B an extension of it, k ;4 X, and k ¢ a,(B).
Prove that B is
a
J-bi-extension of the operator A if and only if
its
Cayley-Neyman transform Kx(B) is a J-bi-extension of the J-isometric operator K),(A).
Hint: Use the definitions of J-bi-extensions of J-symmetric and J-isometric operators, and also 2.Formulae (6.29) and (6.31). 7
Prove that A E St(.Y; ko) if and only if ko 0 ao(A) and K0(A) E St W. Hint: Use 2.Propositionl6.11.1
8
Let A E St(.W; ko). Prove that among its maximal J-symmetric extensions there are operators A E St(.YY; ko) (Azizov).
Hint: Combine the results of Exercises 6, 7, Theorem 2.2 and Corollary 2.3. 9
Prove that a J-symmetric operator A E St(.W'; ko) will be a maximal J-symmetric operator if and only if at least one of the following conditions holds: a) ko E p(A); b) ko E p(A); RA - aot and A- a.t are uniformly definite subspaces of different signs (Azizov [14]).
Hint: Use Corollary 2.3, having first verified that a J-symmetric operator A E St(.YP; ko) is maximal if and only if the J-isometric operator K>,,,(A) E St ,Y is maximal. 10
Prove that every J-symmetric operator A E St(.Y'; ko) admits J-bi-extensions Ao such that one of the points Xo or Xo is regular for them, and the other is a point of
regular type and .RA - k,i,
A - ),,,i
are projectionally complete subspaces
(Azizov).
Hint: Use the result of Exercise 7 and Corollary 2.4. 11
Suppose that A E St(.W; ko) and that it has a maximal J-dissipative extension Ao with Xo E p(Ao). Prove that the operator A has maximal J-dissipative extensions A E St(.YP; ko) with ko E p (A) (Azizov [ 14] ).
Hint: Use the result of Exercise 6, Remark 2.16, Proposition 2.15, and 2.Proposition 6.11. If necessary consider 2. Remark 4.29. 12
Let A and AO be the same as in Exercise 10, ko E p(Ao), Im ko < 0, and let Uo = Kx0(Ao). Prove that the values of the resolvents Ra0(A) of all J-bi-extensions
A of the operator A with ko E p(A) are described at the point ko by the formula R4(A) = (I>,0Uo(UoUo)-` - I)12i Im Xo,
where lx traverses the set of all J-bi-extensions of the operator I I WA - 41; if, in addition, Ao E St(.W; ko), then R,,0(A) _
I)12i Im Xo,
and when AO = Ao the operator A will be J-symmetric (respectively, maximal J-dissipative, J-self-adjoint) if and only if ix, is a J-semi-unitary (respectively, J-bi-non-expansive, J-unitary) operator (Azizov [14]). Hint: Use the results of Exercise 16 on §1, Exercise 6 and 2.Formula (6.29). 13
Let A E St(./P; ko) and Im ko < 0. Prove that A admits extensions A =Ac E St(.M; Xo) with exit into some J-space .W D .W'. Moreover the following conditions are equivlent: a) A admits x-regular ko-standard extensions A = A `;
b) A admits regular ko-standard extensions 4= c) A admits maximal J-dissipative extensions A E St(. Y; ko) with (Azizov [14]).
Hint: Use Theorem 2.18 and the result of Exercise 6.
Xo E p (A )
266 14
5 Theory of Extensions of Isometric and Symmetric Operators Let .w be a J-space, N' +O 1P a J-space, where J = J Q+ 1, and let A be a J-symmetric operator. Prove that A admits Jself-adjoint extensions (cf. Exercise 13).
Hint: Use the scheme A -* JA -. JA O+ - JA -* JA O+ - JA = J JA 15
- JA.
VV= let the (.YYt , .W1 )-decomposition and Let V E St(.YPI, IY2 ), Jo [+] J, [+] V2 and the (.Yr z , .Wz )-decomposition W p = .moo [+] .-?, [+],R2 have the
property V11; = , ,, i = 1, 2. Then the following conditions are equivalent:
a) the operator A has at least one (J,, J2)-unitary extension; b) there is a (J,, J2)-unitary operator W such that WJv= Rv and WJ, = -R,;
c) there is a (J,, J2)-unitary operator U such that UJv=.wand UJ2=.-R2; d) there is a (J,, J2)-isometric operator V, mapping 91F1i on to .1,1.11 and in the Krein space VJ2 [ + ] V, (J i1i )2 there is a dual pair (Y" Y'-) maximal in that space and such that SP± fl VJz= (B) and Y'± fl V,(Ji1))2= (0) (Azizov [14]).
§3
Generalized resolvents of symmetric operators
Before describing generalized resolvents of J-symmetric operators we introduce and prove a number of auxiliary propositions which are, incidentally, of some independent interest. One of the concepts generalizing the 1
concept of an `extension' of operators is that of a `dilatation' of operators. An extension is a process which can take place both within the limits of the given spaces or with emergence from them; but a dilatation takes place necessarily with emergence from the original spaces. Definition 3.1: Let YW; (i = 1, 2, 3) be Hilbert spaces, and let T: ,-'de2 be a bounded operator with JT= .YW1. We call a bounded operator T:.Yy1 O+
O+
Y3 a dilatation of the operator T if T= 11
T;;jj i;= 1,
where T,1 = T and T21, T12, T22 are operators such that T12 T 2 t T21 = 0 when
n = 0, 1, 2, .... If also the Wi are J;-spaces, and if T is a (J1, J2)-unitary operator with it = J, O J3, JJ = J2 O J3, then we call T a (J,, J2)-unitary dilatation of the operator T We say that such a dilatation is x-regular if .W3 = II,Y with x negative squares; we shall call 0-regular dilatations regular.
Remark 3.2:
If under the conditions of Definition 3.1 Tis an extension of the
operator T (this is equivalent to the equality T21 = 0), then, clearly, T is a dilatation of the operator T. However, not every dilatation is an extension. We shall convince ourselves of this later, for example, in Theorem 3.4. Lemma 3.3:
Let T:.-WI -' X2 be a bounded operator with JT =,W,, and
be a dilatation of . let T: -W, 2 1Y3 3 T: .Ye,_O .W'3 O+ J 4 , Y'2 ( W3 O+ tea be a dilatation of
it,
and
let
the operator T.
Then T is a dilation of the operator T By hypothesis the operators T and T admit the follwing matrix represen-
§3 Generalized resolvents of symmetric operators
267
tations relative to the corresponding decompositions: T= II T%JII?J=I,
T=IITJII'J
with
T21=0,
and I
I T3 I I ?= I T33II T 3 j 1
1
=o = 0 (n = 0, 1, 2, ...).
Hence II Tu IIJ=2(II T,J 11 J=2)"II TJ1 IIJ=2 n-I
T12 T22 T21 + T13 T33 131 + T13 k=1
T33T32Tzz ` kT21
= 0 when n =0, 1, 2, ..., and it only remains to use Definition 3.1.
Theorem 3.4: Let T:,W, .W'2 be a bounded operator acting from a J,-space WI into a J2-space i 2 with CA T ='I. Then there is a J3-space W3 and a (J1 (@ J3, J2 O+ J3)-unitary operator T:.O, O+ W3 - -W2 O+ .03 which
is a dilatation of the operator T Let J, - T*J2 T = Jo I J, - T*J2 T I be the polar decomposition of the selfadjoint operator J, - T*JzT, let moo =3j, - T'J2T and let Jo = Jo I .Wo
Then Jo = Jo = Jo 1. We bring into consideration the Jo-space WO' = Q+
0
-.W o', where ,moo
= Yeo and Jo= (
k=0
JAkt, Jak)= Jo (k=0, 1,2,...),
and we define an operator T, :, O+ 06' -'2 O+ 0o by the formula
T1(x;xo,xl,...)=(Tx;IJI-T*J2TI1i2x,xo,X1,...).
(3.1)
By construction T, is a continuous operator with IT, = 'Y, 0,06' and it is (J1 O+ Jo, J2 O+ Jo)-isometric, i.e., T, is a (J, Q+ Jo, J2 O+ Jo)-semi-unitary
operator. The subspace moo is invariant relative to Ti, and therefore T, is a dilatation of the operator T. By virtue of Theorem 2.18 there is a J-space .W
such that the operator T, admits a (J, O J3, J2 O J3 )-unitary extension T: .01 O+ .3 -'2 O 03, where .3 = Wo O+ ' and J3 = JO' O+ J. It only remains to use Lemma 3.3. Corollary 3.5: For an operator T under the conditions of Theorem 3.4 to admit a regular dilatation it is necessary and sufficient that it be a (J,, J2)-binon-expansive operator. El The necessity follows from Exercise 34 on 2.§4. The proof of sufficiency can be carried out using the same scheme as in the proof of Theorem 3.4 taking into account that Jo = I I iJ, - T'J.T and that the operator T, is
268
5 Theory of Extensions of Isometric and Symmetric Operators
(J, O+ J6, J2 O+ J6)-bi-non-expansive. Consequently, by virtue of Theorem 2.18 we can put J = I. 2 Here we shall generalize the concept of a dilatation (see Exercise 4 below) and investigate special classes of holomorphic operator-functions.
Let T(µ) be a function holomorphic in the neighbourhood of 0 with values in a set of continuous operators acting from a space .I', into a Definition 3.6:
(B space .W'2 with VT(,) _ . 1, and let T= II T,, II ?i=1 :. 3 -,02 (9 '3 be a bounded operator with 1'r = 1Y1 O+ J °3. We shall say that the function T(u) 1
is generated by the operator T (or that the operator T generates the function T(µ)) if in some neighbourhood of zero T(µ) = T + µT,2(13 - µT22)- `T21.
Definition 3.7:
In Definition 3.6 suppose the .Yt'i are J,-spaces (i = 1, 2, 3) and
that T is a (J, O+ J3, J2 O+ J3 )-unitary operator. Then we shall say that the
function T(µ) belongs to the class H(MI, .2). If, in addition, .W'3 is a Pontryagin space with x negative squares, then we shall say that T(µ) belongs
to the class fl W',,.Y'2). (In cases where it will not cause confusion, the symbols 01, W2 will be omitted from the designations of classes). Lemma 3.8:
Let T= I I Tii
11 i=
i :.Yf, 0 JP3 - W2 O+ .k3
and
J4
W1
T=IITUII1i=i:.Y,
be bounded operators with #T = .N, O+ .YY3, VT= .-W, O+ .Yr'3 O+ .Yea, and let
T be a dilatation of the operator T Then T and T generate the same operator function T(µ) : Ye,
Ye2.
In accordance with Definition 3.6 we have to verify that T11 + µTlz(I3 - µT22) - `Tzl
= T + It II Tii II i=z(I3 G4 - µ II Tii
II'i=2)
`II
Ti,
II'=2.
(3.2)
The condition 'T is a dilatation of T' is equivalent in the present case to the system of equalities T3 T 33 T3j = 0 (i, j = 2, 3), and therefore II T,iHI]=2(II TiiII'i=2)'II Ti li=z= T12Tz2T21
(k = 0, 1,2).
It only remains to note that in a sufficiently small neighbourhood of zero the equality (3.2) is equivalent to the coincidence of the series k= =0
µkT12T22T2,
and
k=0
µk11 TUIIJ=2(11
TiiII'i.i=2)kII Ti, II' =2.
§3 Generalized resolvents of symmetric operators Lemma 3.9:
269
Let T(µ) :.WP, --+ k2 be an operator function holomorphic in
the neighbourhood of zero. Then there is a space IY3 and an operator T= II T, II ?i=1 : h"1 T
O+
3 such that T(µ) generates the operator
--W2 O+
3
We consider the function, holomorphic in the neighbourhood of zero,
K(µ) = (T(µ) - T(0))µ with µ 5;6 0 and K(O) = T'(0). By Cauchy's formula we have for it, when r > 0 is sufficiently small, K(µ)
tai
r-µ
t1=
2ir
N)
A
K(rfl d(' = 1
K(reI
d4p
(I it
I < r).
µ
J
Let .03 be the Hilbert space of weakly measurable functions x3 (e "0) on [ - a, n] with values in .Y2 and with the square-summable norm IIx3(e'")I12dtip
let T22 be the operator of multiplication in .1Y3 by r-'e-"°, let
T21x, = K(re' )xi when x, E k,, let A
T12x3 = 1
x3 (e' J
d4p when x3 E --W3, and let T = T(O).
Then the operator T = II T, II ?i= I defined on Wi O+ .l'3 is bounded, acts into ./r'2
.IY3, and T(µ)= T(0)+µK(µ)= Tu+AT12(13-µT22)-'T21.
We summarize the results set out. Theorem 3.10: If T(µ) :.)Y, .Y(2 is an operator function holomorphic in the neighbourhood of zero, then T(µ) E H. 11
It suffices to compare Lemma 3.9, Theorem 3.4, and Lemma 3.8.
In this paragraph we introduce a number of results and constructions on the basis of which we shall obtain a criterion for T(µ) to belong to the class II'.
3
Definition 3.11: A function K(X, µ) of two variables, defined on A x A, where A = A* C C, and with values in a set of continuous operators acting in a Hilbert space .# and defined on it, is called a Hermitian kernel if
K(X, µ) = K*(µ, X); a Hermitian kernel K(X, µ) is said to have x negative squares if for arbitrary finite sets (X;); C A, ((;i12 C C, and (x;); C ./P the quadratic form t.l=,
(K(X;, X )x;, xi)Zi i
(3.3)
270
5 Theory of Extensions of Isometric and Symmetric Operators
has not more than x negative squares, and for at least one set has exactly x negative squares. In particular, if x = 0, then K(X, µ) is called a non-negative Hermitian kernel. El Let ,' be a J-space, let F(k) :.0 W' be a function holomorphic in the neighbourhood WF of zero with values in a set of continuous operators with /JF(X) _ ,Y, and let K(X, µ = J[F(X) + F`(a)]/(1 - Xµ) be a Hermitian kernel
having x negative squares. We shall set up by means of the kernel a correspondence between the function F(X) and a certain Pontryagin space II, (F), we introduce a continuous operator P :.° II, (F), and a certain 7r-semi-unitary operator in and we express the function F(X) by means of them. Let X E v(F, let ex be a symbol, and let x>, be a vector in .,Y. We form a linear set S' consisting of formal sums f = Ex ( Y/, eXx>, in which only a finite number
of the xa i;e 6. For elements f and g = EX E
e),g), we define a form
Y/,
If, glo = Ea.µE 111, (K(X, µ)x>., yµ). By hypothesis this form has x negative squares, and if P is degenerate and Yo is it isotropic part, then the completion of the factor-lineal Y/Yo will be a Pontryagin space with x negative squares
(cf. Exercise 6 on 1.§9)-and we denote it by fl,(F). From now on we shall identify elements from 9 with the corresponding elements from £lYo. On the set
1v= j.f If= aE Yj14 e),xxEY, Y xx=9} XE 14
we define an operator V : Vf= EXE 111, XE)IXX- Since
[Vf, V.f]o= >
x
X.µ E Y/,
X.1+ E Y/,
1-Xµ =
xµ
x
p E Y/
X E Y/, L
F(X)x,\,
F(X) + F`(µ)lxx, 1- Xµ J
xµ
J
([F(k)xa,x,,]+ [xa, F(µ)xµ])
[F(X)x),, X E Y/,
\\1 J
1
µ E Y/,
J
µ E Y/, 1- XtL. xµJ
['Z/,
+
kµ
1-Xµ 1
µ E Y/,
I
xa, F(y)x, ]
XX, F(A)x,
X E W,
=[f,flo, Visa 7r-isometric operator in II, (F). Its range of values fit v consists of vectors g = EX E Y/, e,,ya E 9' such that yo = 0 and Eo , a E Y,, y,\/ X = 0. This set is dense in
n, (F). For, if e; x -' eox, ('e),x
--+ 0 and x E ., then, using Exercise 13 on 1. §9, we obtain 0, and therefore any element E,, E Yn exx7, E ' (xo = 0) is
approximated by the elements EXE Y/, eaxa - (et Eo4 XE Y,, x>/X E R v, and since
Y' = II, (F), it follows that ,;? v = rI, (F) also. It follows from the last relation that is non-degenerate, and this implies that V is invertible and hence
(taking 2.Corollary 4.8 into account) that the operators V and V-1 are
§3 Generalized resolvents of symmetric operators
271
continuous. We keep the same notation for the closures of these operators, the second of which will be 7r-semi-unitary. We bring into consideration an operator IF : Ye -p 2' by putting Fx = (1 /, 2)eox, x E M. For any x, y E Y and X E WF we obtain [Fx, Fx]o = [(F(0))Rx, y], [rx, exy]o = (1/,2) [F(0) + F`(X))x, y], and therefore IF is a continuous operator (see Exercise 6 on 1 §9). Consequently there is the operator r': rl (F) .Y and
FT = (F(O))R,
r `c),x = (1 ,,!2) (F` (0) + F(X))x.
(3.4)
Since (see 2.Definition 6.6) the spectrum of the ir-semi-unitary operator V ', with the exception of not more than x eigenvalues, is situated in the disc (t t 15 1), the resolvent (V - XI) - ' exists everywhere in the disc (X I I X I < 1) with the exception, possibly, of not more than x points. Since
(V- XI)(e),x- cox) = Xcox, we have r`cxx-r`eox= XT`(V- XI)- Icox, and so we obtain from (3.4)
F(X) = i(F(0))r+ F`(V+ XI)(V- VI)-'F.
(3.5)
We sum up the foregoing argument:
Let V' be a J-space, let F(X) be an operator function holomorphic in the neighbourhood of zero and taking values in the set of continuous operators defined on ', and let the kernel K(X, µ) = J(F(X) + F`(µ)) (1 - X) have x negative squares. Then: Theorem 3.12:
1) F(X) extends, preserving these properties, on to the disc [X I I X I < 1) with the possible exception of x points; 2) there is a space II,,, a bounded operator r : -+ II,,, and a a-semi-
unitary operator V ' : H - H. such that the equality (3.5) holds for all
X a(V),IXI<1. E Remark 3.13: In Theorem 3.12 the operator V can be regarded as ir-unitary since, by virtue of Theorem 2.18, it admits regular extensions V and Mr=(V-XI)-'I -f r. (V-XI)-`I The proof of the main Theorem 3.16
in this section depends on the following two lemmas, the first of which is a direct consequence of
Theorem 3.12.
-+ r be an operator function holomorphic in the neighbourhood of zero, let the kernel L(X,µ)= J - T*(µ)JT(X)/(1 - Xµ) Corollary 3.14: Let T(X) :
have x negative squares, and let 1 E p(T(O)). Then T(X) E II k. Since 1 E p(T(0)), there is a neighbourhood -V/o of zero such that 1 E p(T(X)) when X E ?lo. We put F(X) = (rzI + a T(X )) (I - TX)-' when X E 40/0. The function F(X) is holomorphic on Rio, and since T*(µ) JT(X) F(X) + F`(µ) _ (I - T(1\)) , 1 - µX - 2 Re a(1- T *(µ)) ' J- 1 1
272
5 Theory of Extensions of Isometric and Symmetric Operators
so when Re a > 0 the function F(X) satisfies the conditions of Theorem 3.12. Let II,,, r and V be the same as in Theorem 3.12, and suppose (see Remark 3.13) that V is a IIx-unitary operator. Since T(X) is expressed in terms of F(X) by the formula
T(X)_(F(X)-czl)(F(X)=al)-' = 1-2 Re a(F(X)+al)-', so, using (3.5) we obtain
T(X) = (F(0) - al) (F(0) + al) + X(24 Re a (F(0) + al)-'P ` V- (1- X (V-' - 2I' (F(0) + al) -'P ` V-'))-' x (2,jRe a F (F(0)) + aI)-1).
We define in the space If O+ 11, an operator T= II Tai II ?i= 1, where
T = T(0)=(F(0)-dI)(F(0)+al)
1,
T2, = 2, Re a (F(0) + aI) -',
T22 = V- 1 - 2I' (F(0) + al) - 1I' ` V-'.
T12=2,Re a(F(0)+a!)-'I"V-',
This operator generates the function T(X), and it is verified immediately (takinn the equality FT = (F(0))R into account) that T`T = T`T = I, i.e., T is a (J + operator. Let To: . , -.-*12 be a continuous operator, with 9'T, =,W,. There will be a x-dimensional space II_ and a (J(, J2)-bi-non-expansive
Lemma 3.15:
operator T= 11 Ti; Hjz j=, acting from the J;-space i = . , O+ II_ into the J2'-space W2=.W2 O+ II_, with J; = J; O+ - I_ (i = 1, 2), and such that T11 = To, if and only if the operators J, - To J2 To and J2 - To J, T* have the
same number x, 5 x of negative eigenvalues (taking multiplicity into account). Under these conditions the operator To is expressible either in the form To= RW, where W is a (J,, J2)-bi-non-expansive (J,, J2)-semi-unitary
operator, and R :.i'2 - W2 is a bounded operator with Ker R J Ker W` and the set [C\(IR U iIR)] fl a(R) is no more than countable, in the form To = W' R', where W' is an operator (J2, J, )-conjugate to a (J2, J,)-bi-non-
expansive (J2, J, )-semi-unitary operator, and R' :.1 -., is a bounded operator with Ker R" D Ker W' and the set a(R) E [C\(1R U iR)] is no more than countable. El Suppose the subspace II_ and the operator T indicated in the Lemma exist. In accordance with 4.Theorem 1.12 the operator T` admits a (J2, J, )polar decomposition T` = V`S, where without loss of generality (otherwise we would consider the operator To instead of To) we can suppose that V is a
(J,, J2 )-semi-unitary (J,, J2)-bi-non-expansive operator, and Sc has the properties of a J2-module indicated in 4.Theorem 1.12. Hence T = S V. Let P;
be an orthoprojector from .'I on to Y;, i = 1, 2. Then To = P2 S VP1 I .%P, .
§3 Generalized resolvents of symmetric operators
273
If
H-
S = II S.! II ?J= 1 :,-W2
--W'2
n-
and
fI_'.V2 1-1V= are the matrix representations of the operators S and V, then To = S11 V11 + S12 V21. Just as in the proof of the implication b) - c) in Theorem 2.18, we prove that the operator V11 I Ker V21 admits a (J1, J2 )-semi-unitary (J1, J2)-bi-non-expansive extension W, and we let P be a J2-orthoprojector from W2 on to Rw. Then VaiII?J=1:.-W1
TO= (S11 P + S11(V11 W` - P) + S12 V21 W`) W = R W. By contruction W has
the properties indicated in the Lemma, and Ker R 3 Ker W`(= RV). We verify that a(R) fl [C\(IR U iIR)] consists of not more than a countable set of points. Since the operators S11(V11 W` - P) and S12 V21 Ware finitedimensional, it is sufficient by virtue of 2.Theorem 2.11 to establish that the set
a(S11P) fl [C\(IR U iIR)] has the property mentioned. In accordance with Corollary 3.5 the J2-selfadjoint operator S, which is J2-bi-non-expansive, admits a regular dilatation S which is a J2"-unitary operator acting in the JtZ = W2 O+ II- O+ fl,, where J2" = J2 O+ - I_ O+ I+. Let Yez = 3 w [+] RVI, and let 9= II S;; II i;= 1 be the matrix representation of the operator S relative to this decomposition of 2 . Since a(S11P) - (0} = a(PS11 P) - (0), and PSP I Mw = S11 (= 911), it is sufficient to verify that the set fl [C\(IR U iR)] is no more that countable. Since the operator S is J2-unitary, it follows that 11 - 921, = 92,,§21. Let Xo be the boundary point of the spectrum of the operator I1 - 92, 1. Then there is a sequence (x,, } E Mw such J2" -space
that 0
S21Szl which is a (I- P)* Jz (I - P)-selfadjoint operator. Since W is a (J1, J2)-bi-non-expansive operator, the J2"-orthogonal complement Iw, to .l?w is a Pontryagin space with x negative squares. Therefore (see 2.Corollary
3.15) a(I, - S21)fl (C\IR) consists of not more than 2x eigenvalues (taking multiplicity into account), and therefore fl [C\ (R U iIR)] consists of a finite number of eigenvalues of finite multiplicity. Thus To = R W1 where R and W have the properties indicated in the Lemma.
Since T is a (J;, J2)-bi-non-expansive operator, there is, by virtue of Corollary 3.5, a Hilbert space fI+ and a (J,, J2)-unitary operator T= II T;; II ? i= 1
acting from the Jspace '1 = .,'1 (1 (II_ O+ fI+) into the
J2-space ie2 =..Y2 O+ (H_ ( f1+ ), where J; = J; O+ ir, and such that IF,, = To. The equalities J1 - To J2 To = T2, rr. T21
and
I_ + I+,
J2 - To J, To = T127r. T12
can be verified immediately. Consequently (see Exercise 14 on 1§9) the
5 Theory of Extensions of Isometric and Symmetric Operators
274
operators J, - 7o J2 To and J2 - To J, To* have respectively x 15 x and x2 < x negative eigenvalues (taking multiplicity into account). We recall that x, and x2 coincide with the number of negative squares of the forms [(I, - ToTo) , ], and [(I2 - To To) , ]2 respectively. Since
I, - To To = II - W`R`RW = W`(I2 - R`R)W, and Ker R 3 Ker W`, and Ker W` is a uniformly positive subspace, it follows that x, coincides with the number of negative squares of the form [(12 - R`R) , ]2. We note also that
from the condition Ker R 3 Ker W` the equality 12 - To T` = I2 - RR` follows.
Let X be a linear operator; we shall denote by the symbol X, the operator e' '`X (SO E FR). It is easy to see that
I2-R,(R,,)`=I2-RR`.
and
Since the set o(R) fl [C\(IR U iIR)] is no more than countable, there is a number po such that 1 E p (R,,). Consequently I2 - R `R = (12 - R,,)` [ (I2 - R,0) ' + (12 - (R,0)`) -' - 12l (12 - R,0),
12-RR`=(12-RPo)[(12-R,,,,)-'+(12-(R,p,)`)-'-12](12-(R,)`), and therefore
12 - R`R = (12 - (Rp)`)(12 - Rw0))-'(I2 - RR`)(I2 - (R,0)`)-'(I2 - R,,,), which implies that the numbers x, and x2 are indeed the same. Conversely, let the operator To be such that each of the operators J, - T, J2 To and J2 - To J, To* has exactly x 1 5 x negative eigenvectors (taking
multiplicity into account). We shall first suppose that dim J", = dim W2. We bring into consideration the operators TV) = To Fn,
where F" =
+1
P1 + rn
P1
(n = 1, 2, ...)
For sufficiently large n the negative spectrum of each of the operators J, - To)*J2 To") and J2 - T(n)J,To") consists of exactly x, (taking multiplicity into account) eigenvalues and these operators are boundedly invertible.
We denote by J;") and j2(n) respectively the unitary parts of the polar decompositions of these operators, we put Pi,, = z (I; J;")), i = 1, 2, and we ± bring into consideration the j;')-spaces it- ,") = .YP; O+ .Ye,, J;") = J; ( j(n) and the operators T(n)
= II
7{n) II2 . . Wi(n) t.J= l .. ,j
Wi(n) 2
where ,f
71
Tin) = Ton)
-I J, -
72(n) = 1
T2(121)
1/2 yn, n) = I J2 - l on) J17pn)*I
7n)* J2 T(n) 1/2, o
o
I 0 J(111)1 JI - P011)*
0
0
J2-
0
0
where Vn is a unitary operator mapping Ye, and to .02 (its existence follows
§3 Generalized resolvents of symmetric operators
275
from the fact that dim .)Pi = dim .IY2) and moreover mapping P,;,,. 1 on to P2.,,.Y('2. It can be verified immediately that T(' is a (J,("), J2("))-unitary operator. Consequently (see Exercise 34 on 2.§4) the operator f,(n)
= I Ti") I?i=1:.%t'1 +
P 1,n..w'1,
where
Tii)= T(i), Tzi)= P1,nTii), Tiz = Tiz)Pl,n, and f = P1,nTzi) P1, n, - Ii,n, J2 O+ - Ii,n)-bi-non-expansive operator. The operators P;-" converge in norm to P;,o when n -' oo, where P;;o = (I - P,,o - J;, o) (i = 1, 2), is a (Jl
i P1,o is the orthoprojector on to Ker(J1 - T'J2To), P2,o is the orthoprojector on to Ker(J2 - To J1 To), and Jl,o and J2,o are the partially isometric parts of the polar decompositions of the operators J1 - To J2 To and J - To J1 To* respectively. We choose Vn so that the operators VnPi,n converge in norm to VPI;o, where V: Plo 1 -' Pzo.W2 is an isometry. Then the operator Pj,oa)l
T= II TjiII?.i=:,Wl(@ with T= To, T21 = (Pl.o(T0J2T0 -
T12 = (P2.o(To J1 To - .12)P2,o)1/2 VP1;o, J1)P170)1/2'
T22
= T 'P1.oToJ2T12
will indeed be the required (Ji, J2)-bi-non-expansive operator
(J;= J; G) -Ii,o,i=1,2). We now suppose that dim .W1 ;e dim a2, and that dimensional Hilbert space with dim ,4
,
'
is an infinite-
max (dim .1, dim ,Yt'2 l . We consider
an operator TV acting from the (Jl O+ I)-space k1 O+ ' into the (J2 O+ I)space .2 O+ . and which is an extension of the operator T, and moveover TV I .W = I. The operator T1S satisfies the same conditions as the operator To,
and dim(.1 O+ .) = dim(,Y2 0 .0), and therefore the proposition proved above holds for it. It only remains to use the result of Exercise 34 on 2.§4.
Theorem 3.16: Let T(X) = Z' O X' T1 bean operator function holomorphic in the neighbourhood of zero. Then the following assertions are equivalent:
a) T(X) E rl';
b) the negative parts of the spectra of each of the operators J;n) - T(n)* JZn)T(n) consist of not more than x negative eigenvalues, and the
negative parts of the spectra of the operators Jf O) - T (O)* J1°) T(O) and J)°) - T (O) J1 T (O)* consist of an equal number x 1 <, x of negative eigenvalues
(taking multiplicity into account), where J°r = Jk, Jkn) = Zk=o O+ Jk, and T'n) = II T1 II"i=o are Toeplitz matrices acting from the J(n)-space ,y,n) into
5 Theory of Extensions of Isometric and Symmetric Operators
276
the J2(")-space .YPZand Tij = 0 when i > j, T;; = Tj_ 1 when i <, j, k(n)_ yek
.%Pko)_yek,
L
...
i,j=0,1,2,...,n;
(DWkg J
n+l
c) the kernel [J1 - T*(µ)J2T(X)]/(1 -µX) has not more than x negative squares, and the negative parts of the spectra of each of the operators J1 - To J2 To and J2 - To J1 To* consist of x, < x2 eigenvalues (taking multiplicity into account).
a) - b). Let T(X) E rIx, i.e., there is a Pontryagin rx-space rrx = rI_ O+ rI+ with dim rI_ = x < oo and a (J1 O+ rx, J2 O+ rx)-unitary O + Hx - M22 operator T= I I T;jjj 1 j= 1 : rIx such that in the neighbour(
1
hood of zero T(X = T11 + X T12 (Ix - X T22) _' T21. Hence
To =T11,T;=T12Tzz1T21) i=1,2,..., and
Jln)- T(n)*JZn)T(n)= 7'(n)*J(n)T(n), where T<") = II
Tij II ;'j=0: Jyj 0 ...
o
,
rrx
o ... o
n+1
rrx
n+1
is a diagonal matrix with T, = T22T21 (i = 0, 1, 2, ...), and J(") = I JU
0:1Ix
...
rIx
n+1
rrx O+ ... O+ rIx, Jij = rx, n+1
i, j=0, 1,2,..., n; n=0, 1,2,.... Since the selfadjoint operators J(") have not more than x negative eigenvalues, it follows from Exercise 14 on 1.§9 that the operators J;") - T(")* J,(")7{") also have not more than x negative eigenvalues (n = 0, 1, 2,. ..).
The validity of the assertion that the operators
Ji0)
- T(0)* J2 T(0) and
J2(0) - T(0)J;0)T(0)* have the same number of negative eigenvalues follows from Lemma 3.15 and Exercise 34 on 2.§4. b) - c). Since o( J;") - T(")* Jz")T(")) fl (- oo, 0) consists of not more than x negative eigenvalues (taking multiplicity into account), and this is equivalent to the fact that the forms ((.Iin) - T(n)* JZn)T(n))X(n), x("))
have in, O+ ...
O+ ,Ye'
+rl
not more than x negative squares, it follows that for arbitrary finite sets of vectors (x1} " C .,Y1 and arbitrary complex numbers (Xj] 1 the form
((Jin) - T(n) J2nT(n))x,(.s), x,j'4), r, s, p, q
§3 Generalized resolvents of symmetric operators
277
where
O
X; s) = (Xr, X5Xr, ..., X s Xr) E .,Y1
... O
.
1,
(n + I) times
s= 1,2,...,t; r= 1,2,...,m, has no more than x negative squares. This form can be rewritten as
((1_X)(JiXrXP)i
1
r.s.P,q
1 - X5X
+Z
(1 - Xs
J2T x , Xr)1
i.J=o
((1 -
V= qTJ*) J2 \i
r,,p,q
0
xsT,/) Xr, XP)
1
1 - XSXq XP)1 1\
I - X, 1q
r, s, p, q
n-ljn-1
n
s
+
i.J=0
^9 (_1 - XsXq
(X lT!)J2(X'sTi)Xr, XP
To prove our assertion it suffices to verify that the second sum over r, s, p, q tends to zero as n oo if we take the (Xi) i in a sufficiently small neighbour-
hood of zero. Without loss of generality we shall suppose that
II Xr II = 1
(r=1,2,...,m),andIX <e(s=1,2,...,t)ande'12II TTII <, 1(j=1,2,...) (the latter is possible because T(X) is holomorphic in the neighbourhood of zero). Since there is only a finite number of terms in r, s, p, q, we shall verify that each of them tends to zero as n oo; indeed, n
X'nXn(J1Xr, XP)1 + Z (Xs
Ti)Xr, XP)1
i,j = o I - XsXq
2n+
S
n
Z
£2n+(j-3i)/2
i''=o
0
1 - c2
as n - oo.
c) - a). In accordance with Lemma 3.15 we can suppose without loss of generality that To = R W, where R and W have the properties indicated in that lemma. Let tp E [0, 2a) be a number such that 1 E p(R,,). We consider the function which is holomorphic in the neighbourhood of zero. This function satisfies the conditions of Lemma 3.14 and therefore R,(X) E II"(,'2), i.e., there is a aX-space II,, and a (J2 O+ ak)-unitary operator R = II Rii. vll ?i= l :.f(2 G IIX --
ED n.
such that in the neighbourhood of zero R,(X) = R11,,, + XR12,,,(I, - >,R22.,)-IR21,,.
5 Theory of Extensions of Isometric and Symmetric Operators
278
Hence,
T(X) = R,(X)W-,= R11.,W-.c+ XR12,,,(I"- XR22,,)-'R21,,W and therefore T(X) is generated by the matrix T' = II T11II i= 1:.W1 on"-..YP2 orlk, where
Tit = R11,,. W-., Tie= R12,,, T21 = R21, W-,, T22= R22,,,. By construction T' is a (J1 O+ r", J2 O+ r" )-semi-unitary (J1 O+ r", J2 O+ r" )-
bi-non-expansive operator. By virtue of Corollary 3.5 it admits regular dilatations, and it follows from Lemma 3.8 that each of these dilatations generates the function T(X), and therefore T(X) E II". Now let W 1 = W2 = . be a J-space, let T(µ) :.W -p ,W be an operatorfunction holomorphic in the neighbourhood of zero, let 4'o be a Jo-space, and let T= II T, II ?r=1 : ,Y e ,Yo - W 0 o be a J-unitary operator generating the function T(µ), where J= J+ Jo. Definition 3.17: The J-unitary operator T= 11 T1Ij i1= 1 will be called simple if the operators T and T22 have no common completely invariant subspaces.
Theorem 3.18: Every operator function T(#): -.71' holomorphic at zero is generated by a certain simple J-unitary operator T = 11 T.i II?i=1 :.(0
-.R 0 .7(o.
Moreover, if T(µ) E IT, then a Pontryagin space with not more than x negative squares can be taken as ./1'o.
Let .73 be the J3-space which exists by virtue of Lemma 3.9, and let T' = II T,i 11 1 i=1 : -W 0 -Y3 - . W 0 3 be the (J (@ J3)-unitary operator generating the function T(µ) (if T(µ) E IT, then we take as '3 a Pontryagin space with not more than x negative squares). The subspace .1P" _ C Lin ( T' "01
is completely invariant relative to the operator T', it contains
./P, and the operator T' is a dilatation of the operator T" = T' I .7(", and therefore (see Lemma 3.8) the operator T" generates the same function T(µ), and moreover T" is an isometric operator in the indefinite metric induced by
the (J (@ J3)-metric in 7". Let P be the orthoprojector from .0" on to its isotropic part. Then the space ,Yo' = (I - P).,Y" contains . and is a nondegenerate (J O+ G)-space :.(o = 0 W', and the operator T" is a (J (@ G)unitary dilatation of the operator To = (I - P) T" I.7(6 and therefore, again by virtue of Lemma 3.8, To generates the same function T(µ). As the space o we take the completion of ,Y' relative to the norm (I G I x, x)1 /2 (cf.. I .Proposition
6.14); as Jo we take the extension of the isometric part of the polar decomposition of the operator G on to .-Wo; and as T= 11 T, Il zj=1 we take the extension by continuity of To on to .7 0 YW o. It is clear that the operator T
§3 Generalized resolvents of symmetric operators
279
generates the function T(µ) and is f-unitary, where J= J OB Jo, and o = C Lin ( T".-*'] We now verify that T is simple. Let . be %r O+ the common completely invariant subspace of the operators T and T22. Then Y'I1]
is completely invariant relative to T and .X C q'111, and therefore
C Lin(T"k)°°_ C Y I1], i.e., 91111 =.Ye O+ .Yo and therefore £_ (0). We note that if .W'3 is a Pontryagin space with not more than x negative squares, then °o also has this property.
Definition 3.19: An operator-function R;, holomorphic in a neighbourhood of the point k is called a generalized resolvent of a J-symmetric operator A acting in a J-space . if there is a J1-space .1 such that the J Hermitian operator A in the f-space O+ .,Y1 (where j = J (D J1) admits a fselfadjoint extension A with ko E p(A) and R1, = P(, T- XI)-' I , where P is 4
='
the orthoprojector from ie- on to. We call such a function R1, a x-regular generalized resolvent if some Pontryagin space with x < oo negative squares can be taken as .,Y,, and a regular generalized resolvent if x = 0. Definition 3.20: We denote by the symbol II>, (A) the set of all operatorfunctions F(X) :.W' .W, holomorphic in a neighbourhood of the point Xo, which can be expressed in the form F(k) = F1 1
where F(X) = I
- xo F( _ kok ko k 12\I J1)-unitary
kok - k0
+kok -
F,i II ?i=1
\
Fzz l
Fz 1,
extension of the operator
is a (J O+
IA.x,, in the (J (@ J1)-space f O+ . 1; .01 is a certain J1-space; if .1P1 is a Pontryagin space with x negative squares, then we denote this set by II
'A _
rIL(A)
It follows immediately from the definition of the set 11x.(A) that for every k from a certain neighbourhood of the point X0 the function F(X) (EIIao(A )) is a J-bi-extension of the operator IA,XO and the operator-function
G(µ) = F(I )'°I z(1
\
ko - Xoµ
µ) I E H
l
is holomorphic in a neighbourhood of µ = 0. Conversely, suppose that for every k from a certain neighbourhood of the point Xo the operator-function F(X), holomorphic in this neighbourhood, is a J-bi-extension of the operator IA,a,,. Then it belongs to the set II1,,,(A ). For, in accordance with Theorem 3.10 G(a) E H and therefore kok - k0
F(k)=F11+kok-
.0
F
1z
- x0 Fzz 1 I I1 _ ko kkok-xoJ
1
Fz1 ,
and moreover by virtue of Theorem 3.18 we can regard the (J O+ J1)-unitary operator F= II F, II Z;=1 :.k' O+ .k1 -* .Y( O+ .,W1 as simple. (We note that we can take H, as.W1 if and only if G(µ) E H".) We verify that Fis an extension of
5 Theory of Extensions of Isometric and Symmetric Operators
280
the operatorlA,a or, what is the same thing, that F21x = 0 when x E WA - i. Suppose this is not so. Then there is a vector xo E MA - j,r such that F21x0 0, and therefore Y' = C LinI FX"J °m n w'1 * (0) and it is a completely invariant subspace relative to F and F22-and we have obtained a contradiction of the fact that the operator F is simple. Thus we have proved
Let F(X) be a function holomorphic in a neighbourhood of the point X0. Then F(X) E IIx, (A) if and only if, for every X in a certain
Theorem 3.21:
neighbourhood of the point Xo, F(X) is a J-bi-extension of the operator IA,X,. Moreover the conditions (F(X) E Hx'o (A )) and j G(µ) = FI' X01 z(1
are equivalent. The operator F be chosen to be simple.
µ))
E
n'}
F;; i;=1 generating the function F(X) can
Let a J-symmetric operator A E St(,W; X0) with Im Xo < 0, let Ao E St(, ,Y; X0) be a J-bi-extension of the operator A such that Xo E p(Ao) (on the existence of such an Ao see Exercise 10 on §2), and let Uo = K1,,,(A). 5
Theorem 3.22:
The relations R>, = (Ta - XI)-' and Ta=(X0-Xo)[F(X)Uo(UoUo)-i-I]-1+X I
describe all the generalized resolvents of the J-symmetric operator A in the neighbourhood of the point Xo when F(X) traverses the set In the neighbourhood of the point Xo we have R"
= El
[(Tµ-µ1)-1l`.
Let A be a J-selfadjoint extension of the operator A acting in the J-space w' O+ -Y,, and let Xo E p(A). It follows from Exercise 12 on §2 that (A - X01)-' =
1
- x0
[4.00(17C0170-1
- 1],
where U0 is a J bi-extension of the operator U = K,,o (A) coinciding on W' with U01 and Uo I, = (X0/4)11i and !>,, is the corresponding J-bi-extension of the operator IA,XO. In accordance with Hilbert's theorem for the resolvent R,,
1
-
P I - X_
Ilo (1 >,,,Uo(UoUo)-1 - I
Jof
[i U0(UoUO)-'I] I .i
Let Ik,,= II Fiijj?i=I be the matrix representation the operator operator U= has the matrix representation U= 11 U1ijj?i=1, where
Un = FnUo(UoUo)-', and Uiz>'o F,2, x0
.
Then the
i= 1,2.
§3 Generalized resolvents of symmetric operators
281
Therefore in a neighbourhood of the point Xo Hilbert's identity given above can he rewritten in the form r /F(X)Uo(UoUo) I F(X)Uo(UoUo) Xo -
where
Xo - xo
O
\
L
F(X) = F1 1
Xo X - X0
-1
Xo X - X0
+Xo X - Xo F12
IXo X - o
Fzz
Fzi
is an operator-function holomorphic in a neighbourhood of the point X0 which is, for every X from this neighbourhood, a J-bi-extension of the operator IA,X,, i.e., F(X) E II,,,,(A) by virtue of Theorem 3.21. In this same neighbourhood of
the point ko the operators F(X)Uo(UoUo)-' are J-bi-extensions of the 1 E o, (U) that where Rs=(T,-XI) ', Hence, I]-1 + X01. Tx _ (Xo - Xo)[F(X)Uo(UoUo)-' Conversely, let F(X)EIh,,,(A). Then F(X)Uo(UoUo)-` is an operatorfunction holomorphic in a neighbourhood of the point Xo which, by virtue of for all X from Theorem 3.21 is a J-bi-extension of the operator U= this neighbourhood. Therefore
operator U= K,,, (A ), and therefore it follows from
IEa,(F(X)Uo(UoUo)-')
/ T(µ) =
z
F( 1 X0 '
Uo(UoUo)-'
(l
X0-1,A))
is a function holomorphic in a neighbourhood of the point-,u = 0, and therefore '1 and a simple J-unitary operator there is a Ji-space
U=
U;;11 ;=, :
0
, --
.
0 .W'i, with J= J 0 Ji, such that U gener-
ates the function T(µ). It follows from the result given later in Exercise 11 that
we can take as U an extension of the operator U. Therefore by virtue of Exercise 16 on §1 U=I,,oUo(UoUo)-1, where Uo is a J-bi-extension of the operator U coinciding on r with Uo, and Uo I W', = Oo/Xo)h, and 1),0 is the corresponding J-bi-extension of the operator 1A.ao. Since U C Uand 1 E ac(U),
it follows, because the operator U is simple, that 1 E oc(U), i.e., there is a J-selfadjoint operator A such that U = Kx,,(A ). It follows from Exercise 15 on II§6 that A C A, and from the proof of the first theorem that
R,,=(T,,- XI)-', where
Ta=(Xo-&) T Q
I
+ XI
= (Xo- Xo)[F(X)Uo(UoUo)-' - I]-' + Xol. It
then follows from the definition of the generalized resolvent that
R, = (R,,)` = [(T;, - µl)-']` in a neighbourhood of the point Xo. Corollary 3.23: Under the conditions of Theorem 3.22 let A0 be a maximal generates J-dissipative J-symmetric operator. Then the function F(X) E a x-regular generalized resolvent, and conversely, every x -regular generalized resolvent is generated by some function from nx',,(A).
5 Theory of Extensions of Isometric and Symmetric Operators
282
Let the function F(X) E HZ,/(A ). Then
FIIX 12(1-µ))U0E H',
T
and the operator U appearing in the second part of the proof of Theorem 3.22
can be chosen to be a x-regular extension of an operator U= Therefore the generalized resolvent constructed there will be x-regular.
Conversely, let R), be a x-regular generalized resolvent generated by a x-regular extension of the operator A. Then U = Ka (A) = where U0 is a J-bi-non-expansive i-semi-unitary extension of the operator U coinciding with Uo on .W, and Uox = (Xo/Xo)x when xE I ,,. Consequently I>,0UoUo is a i-bi-non-expansive extension of the operator IA,ao and therefore, by virtue of Corollary 3.5 and Lemma 3.8, the function generated by this operator belongs to III (A ). Corollary 3.24: Under the conditions of Corollary 3.23 let J = I, x = 0. Then the function (TX - XI)-' admits a holomorphic extension from the neighbour-
hood of X0 on to C-. On C+ we put (Tx- AI) -' = [(TX- XI)-']`. This extension coincides in C+ U C- with the corresponding regular generalized resolvent of the operator A. By virtue of the result of Exercise 12 on §2
Ta=(X0-Xo)(F(X)Uo-I)-'+X01, and by virtue of Corollary 3.23 we can suppose that F(X) E i.e., there is a Hilbert .4" space and a unitary extension (@.;99, of the operator IAA, such that F= II FjiHI +i= I :,-W O+ X F(X)
= Fi +
.o
Xo
_Xo i - Xo
>-1\0 X - 1\o
F 1z I
X0
X - 1\0
\-i
Fzz
Fz i
Since
II Fzz II < and
),o
X - X0
Xo
X - .o
< 1,
it follows that F(X) admits a holomorphic extension from the neighbourhood of the point Xo on to C-. Moreover the extenson of the function z
G(µ) = F( X01 0 _ 01 E IV( 4 - Xoµ
< 1, and therefore II F(X)II < 1),
which, by virtue of 2.Remark 6.13, implies that the operators Ta are maximally dissipative for all X E C-. Therefore C- C p(TX) (see 2. Lemma 2.8),
and therefore (T,, - XI)-' admits holomorphic extension from the vicinity of
the point Xo on to C-. On C+ we put (T,- XI)-' = [(Tx- XI)-']`. On the other hand, the regular generalized resolvent R,, of the operator A
§3 Generalized resolvents of symmetric operators
283
defined on C U C- does coincide in a neighbourhood of the point X0 with the
function (T,\-XI)-', and therefore R a = (T, - XI)-' when X E C- and µEC+.
Exercises and problems 1
Let.YY, and .Yt'2 be Hilbert spaces, and let T:"Y3 Yj C'r= JJt°,. In [XXIII] an operator T:.Y( O+
.JP2 be a bounded operator with .Yt', O+ 'Y3 with C/T=.yY, ED ,k3
is called a dilatation of the operator T if P,T"I .yt', = T" (n = 0, 1, 2,3 , ...), where P, is the orthoprojector from YY, O+ .Y'3 on to .,Yi. Prove that this definition and Definition 3.1 are equivalent when .Y'i =.YW2. 2
Let T and T be the same operators as in Exercise 1. Prove that the operator Twill
be a dilatation of the operator T if and only if P, (T - XI)-" ,W, = (T- XI)-", n = 1, 2, ..., X E p(T) fl p(T) (see [XXIII] ). 3
Suppose that F(X) is a function holomorphic in a neighbourhood of zero with values in a set of continuous operators acting in a J-space Jr', and that there are: a space II, with x negative squares; a continuous operator r :.YP - 11,; a ir-semiunitary operator V -' : II, - II,; and a J-selfadjoint operator S with !Ys = 'Y; such that F(X) = iS+1'`(V+ X1)(V- xI)-'I', I X I < 1, X a(V). Prove that the kernel K(X,µ)= J[F(X)+ F`(µ)] (1 - Xµ) has not more than x negative squares (cf. M. Krein and Langer [4] ).
4
Let T(X) = To: .01 .V2 be an operator-function holomorphic in a neighbourhood of zero. Prove that an operator T :.yt', O+ 'YO - 2 + . Yo will generate the function T(X) if and only if it is a dilatation of the operator To (Azizov).
5
Prove that an operator To admits a x-regular dilatation if and only if To is a (J,, J2)-bi-non-expansive operator (Azizov [14]). Hint: Use the equivalence of assertions a) and b) in Theorem 3.16 and the result of Exercise 4.
6
Give an example of an operator-function T(X) :.yP, - JI2i holomorphic in a neighbourhood of zero, such that the kernel [J, - T*(µ)J2T(X)]/(I - µX) has x negative squares but nevertheless T(X) ri'.
Hint: Put T(X) - To, where To is a (J,, J2)-semi-unitary operator which is not (J,, J2)-bi-non-expansive, and use either the results of Exercises 4 and 5, or the equivalence of assertions a) and b) in Theorem 3.16. 7
Prove that under the conditions of Theorem 3.16 the function T`(X) E II'. Hint: Verify that if the operator T generates the function T(X), then T` generates the function T`(X).
8
Let T(µ):.;o , -..W2 be an operator-function holomorphic in a neighbourhood of the point µ = 0 with values in a set of continuous operators (i T(,) _ .Y',) acting from the J,-space.YY', into the J2-space .W2. Prove that the following assertions are equivalent: a) the kernels
J,-T*(µ)J2T(X) 1 -µX non-negative;
and
J2-T(X)J, T*(µ)
1 -µA
284
5 Theory of Extensions of Isometric and Symmetric Operators b) in a neighbourhood of zero the T(X) are (J,, J2)-bi-non-expansive and the function w' (T(>,)) = T, (X) admits holomorphic extension on to the disc [XI IXI<1)and,IT,(X)II<, 1; Arov[1]). c) the kernel [I, - T,(µ)T,(X)]/l -µX is non-negative O+ + W' is a (J,,J2)-unitary Hint: Verify that if T= 11 T,iII 2 V. operator (J; = J; + I, i = 1, 2) generating the function T(X), then w' (T) generates T, (X), and conversely.
9
Suppose that under the conditions of Theorem 3.16 J; = I;, 1 = 1, 2, and that T(µ) satisfies at least one of the assertions a)-c) of this theorem. Prove that then there is a a,-space II and a (I, v -,, 12 0 Ir,)-unitary operator T= II Ti; II j;_, :./P, O II, - .1P2 b II, generating T(µ) such that T22 has no neutral eigenvalues (Azizov). Hint: Use Theorem 3.18.
10
Under the conditions of Problem 9 suppose J, = I,, J2 = 12, and T(µ) E II,. Prove that then T(µ) admits holomorphic extension on to the disc ( X I I X I < 1) with the
exception, possibly of not more than x points which will be the poles of this extension (M. Krein and Langer [6] ).
Hint: Use the result of Problem 9 and prove that the function T(µ) can be expressed in the form of a product T(µ) = T, (µ)T2(µ), where T, (µ) E II°, and T, (µ) E II' and is generated by a negative 7r,-space II, with dim n, = x. 11
Let T(µ) be an operator-function, holomorphic in a neighbourhood of the point µ = 0, which is a J-bi-extension of a J-isometric operator U for every µ from a certain neighbourhood of the point µ = 0. Prove that among the simple J-unitary operators generating the function T(µ) there is at least one which is an extension of the operator U (Azizov).
Hint: Use the methods of proof of Lemma 3.9, Theorem 3.4 and Lemma 3.8.
Remarks and bibliographic indications on chapter V The definition of the Potapov-Ginzburg transformation in so general a form was introduced essentially by Shmul'yan [5] (cf. Arov [1]). The traditional PG-transformation when A', = ,W'2, J, = J2 was first introduced and studied by Potapov [1] and later Ginzburg [1] continued the investiga§1
tion. The works of I. Iokhvidov [14], [18], E. Iokhvidov and I. Iokhvidov [1], Shmul'yan [4], [5], Ritzner [4] are devoted to the PG-transformation. The exposition in the text, sometimes new in method, is due to Azizov, although almost all the propositions can be found in parts in the works of the authors just mentioned. We mention, in particular, that Theorem 1.14 was obtained independently and simultaneously by Azizov, E. Iokhvidov and I. Iokvidov [1] and Ritsner [4]. §2 The problems of J-isometric extensions of J-isometric operators in the finite-dimensional case were taken up by Potapov [ 1 ] , and in the case of n. by
1. Iokhvidov [15], [16] (see [XIV]; later M. Krein and Shmul'yan [3], E. lokhvidov [2], [4], [5], Azizov [12], [14] studied these problems for operators acting in Krein spaces. In the work of M. Krein and Shmul'yan [3] Jr and .fit, are regular subspaces, and in the articles of E. Iokhvidov [2], [4], [5]
dim V I1] < oo and dim
oo. The latter built his investigations on
Remarks and bibliographical indications on Chapter 5
285
systematic use of PG-transformations (see, e.g., Exercises 5, 6 on §1). With due regard to the above remarks the remaining results in this section and the method used are due to Azizov. §3.1 The concept of a dilatation on operators acting from one space into
another was first,
it
appears, carried over and applied by Azizov
[14]
(Definition 3.1). The remaining results of this paragraph in the case when the operator acts in a single space are well-known (see Davis [1], Davis and Foras [1], Saktinovich [1], Kuzhel' [8], [9]. In our approach these results were proved by Azizov [14], and Corollary 3.5 we find, essentially in Arov [ 1 ] . §3.2 We find the functions T(µ) in Definition 3.6 in Arov [1], for example, where they are called `transfer functions'. Lemma 3.8 and Theorem 3.10 were proved by Azizov [14], and Lemma 3.5 we find, essentially, in Arov [ 1 ] . §3.3 The investigations in this paragraph, carried out by Azizov [14], were stimulated by a series of papers by M. Krein and Langer [3]-[6]. In such a general formulation for Krein spaces neither Theorem 3.12 nor Lemma 3.14 are given by these authors, but our exposition follows their basic ideas, and often repeats their arguments word for word. Lemma 3.15 is due to Azizov. Theorem 3.16 is also due to Azizov [14]. Partially and in particular cases it is encountered in other authors. We single out the series of papers mentioned above of Krein and Langer, and Arov's article, which initiated the appearance of this result. Theorem 3.18 is due to Azizov [14]. §3.4 The set H>. (A) was introduced into the investigation by Azizov. Theorem 3.21 is also due to him (see Azizov [14]). §3.5
Theorem 3.22 is due to Azizov [14]. The form of writing the
generalized resolvent Rx, = (Tx - XI) - ' was first introduced by A Strauss [1].
Moreover we have used M. Krein's idea (see [I]) of all the generalized resolvents by means of a fixed extension of the operator. For more detail about
other approaches to the description of generalized resolvents and historical information see [IV] and also Etkin [1], [2]. Corollary 3.23 is due to Azizov [14], and Corollary 3.24 to A. Shtraus [1]; we point out that the proof in the text, due to Azizov, is not based on A. Shtraus's result.
REFERENCES
1:
Monographs, Textbooks, Lectures, Survey Articles [I] Akhiezer, N. I., and Glazman, I. M., Theory of linear operators in Hilbert space, 3rd Ed., Vishcha Shkola, Khar'kov, 1978. (English translation, Pitman. London, 1981.) [II] Ando, T., Linear Operators on Krein Spaces, Sapporo, Japan, 1979.
[III] Azizov, T. Ya., and lokhvidov, I. Linear operators in Hilbert spaces with a G-metric. (Russ.), Uspekhi Mat. Nauk, 1971, 26(4). [IV] Azizov, T. Ya., and Iokhvidov, I. Linear operators in spaces with an indefinite
metric (Russ.). In the book Mathematical Analysis, 17 (Itogi nauka
i
tekhniki), 113-205, Moscow, 1979. [V] Bognar, J. Indefinite Inner-product Spaces, Springer, Berlin, 1974.
[VI] Daletskiy, Yu. L., and Krein, M. G., Stable Solutions of Differential Equations in a Banach Space (Russ.), Nauka, Moscow, 1970. [VII] Dunford, N., and Schwarz, J. T. Linear Operators. General Theory, WileyInterscience, New York, 1958.
[VIII] Ginzburg, Yu. P., and lokhvidov,
I. S.,
Investigation by geometry of
infinite-dimensional spaces with a bilinear metric (Russ.), Uspekhi Mat. Nauk, 1962, 17(4), 3-56. [IX] Glicksberg, I. L., A further generalization of the Kakutani fixed-point
theorem, with application to Nash equilibrium points, Proc. Amer. Math. Soc., 1952, 3.
[X] Gokhberg, I. Ts., and Krein, M. G., Basic propositions on defect numbers, root numbers, and indices of linear operators (Russ.), Uspekhi Mat. Nauk, 1957, 12(2), 43-118. English translation. Amer. Math. Soc. Translations (2), 13, 1960. 185-266.
[XI] Gokhberg, I. Ts., and Krein, M. G., Introduction to the theory of linear non-selfadjoint operators in Hilbert space (Russ.), Nauka, Moscow, 1965. [XII] Greenleaf, F. P., Invariant Means on Topological Groups and their Applications, Van Nostrand-Reinhold Co., New York, 1970.
[XIII] Hille, E., and Phillips, R. S., Functional analysis and semi-groups. American Math. Soc. Coll. Publications, 31, 1957. [XIV] Iokhvidov, I. S., and Krein, M. G., Spectral theory of operators in spaces with
an indefinite metric, I (Russ.), Trudy Moskov, Mat. Obshch., 1965, 5, 367-432; Amer. Math. Soc. Translations (2), 1960, 105-75. [XV] Iokhvidov, I. S., and Krein, M. G., Spectral theory of operators in spaces with
an indefinite metric, II (Russ.), Trudy Moskov, Mat. Obshch, 1969, 8, 413-96; Amer. Math. Soc. Translations (2), 34, 283-373. 286
References
287
[XVI] Iokhvidov, I. S., Krein, M. G., and Langer, H., Introduction to the Spectral Theory of Operators in Spaces with an Indefinite Metic, Berlin, AkademieVerlag, 1982.
[XVII] Krein, M. G., Introduction to the geometry of indefinite J-spaces and the theory of operators in those spaces (Russ.). In the book Second Summer Mathematical School, I, Kiev, 1965, 15-92. [XVIII] Krein, S. G., Linear Differential Equations in a Banach Space (Russ.), Nauka, Moscow, 1968. [XIX] Krein, S. G., Linear Differential Equations in a Banach Space (Russ.), Nauka, Moscow, 1971.
[XX] Langer, H., Spectral functions of definitizable operators in Krein spaces, Lecture Notes in Mathematics, 1982, No. 948, 1-46. [XXI] Nagy, K. L., State Vector Spaces with Indefinite Metric in Quantum Field Theory, Noordhoff, Groningen, 1966. [XXII] Riesz, F. and Szekefaldi-Nagy, B., Functional Analysis, Ungar, New York, 1955.
[XXII] Szekefaldi-Nagy, B., and Foias, C., Harmonic Analysis of Operators in a Hilbert Space Akad. Kaido, Budapest, 1967. 2
Articles in Journals Akopyan, R. V.
[1] On the formula for traces in the theory of perturbations for J-non-negative operators (Russ.), Dokladi Akad. Nauk, Arm. SSR, 1973, 57(4), 193-9.
[2] On the theory of perturbations of a J-positive operator (Russ.), Functional analysis and its applications, 1975, 9(2), 61-2. [3] On the theory of the spectral function of a J-non-negative operator, Izvestiaya Akad. Nauk, Arm. SSR. Series phys.-mat. nauk. 1978, 13(2). [4] On the formula for traces of J-non-negative operators with nuclear, perturbations. Doklady Akad. Nauk, Arm. SSR, 1983, 77 (5).
Aronszajn, N. [1] Quadratic forms on vector spaces, Proceedings Internat. Sympos. Linear Spaces, Jerusalem, 1960, Acad. Press, Oxford, 29-87.
Arov, D. E. [1]
Passive linear stationary dynamic systems (Russ.), Siber. Mat. Zhurnal, 1979, 20(2), 211-28.
Askerov, N. K., Krein, S. G., and Langer, G. I. [ 1 ] A problem on oscillations of a viscous liquid and the operator equations connected with it (Russ.), Functional analysis and its applications, 1968, 2(1), 21-31. Azizov, T. Ya.
[1] Criterion for non-degeneracy of a closed linear envelope of root vectors of J-disipative completely continuous operators in a Pontryagin space flt, Collection of Candidates' papers, Math. Faculty Voronesh Univ, 1971, ed. 1, 1-5. (Russ.). [2] On the spectra of certain classes of operators in Hilbert space. (Russ.), Mat. Zarnetki, 1971, 9(3), 303-310.
288
References
[3] On the decomposition of a normed space into the direct or approximate sum of special subspaces (Russ.), Trudy N. I. I. Mat, Voronezh Univ, 1971, ed. 3, 1-3. [4] Invariant subspaces and criteria for completeness of a system of root vectors of J-dissipative operators in a Pontryagin space II, (Russ.), Doklady Akad. Nauk, USSR, 1971, 200(5), 1015-17. (English translation, Soviet Math. Dokl., 12(7), 1513-14.)
[5] Dissipative operators in a space with an indefinite metric (Russ.), Candidate's dissert., Voronezh, 1972.
[6] On invariant subspaces of a family of commutating operators in a Pontryagin space H, (Russ.), Functional analysis and its applications, 1972, 6(3), 59-60. [7] On completely continuous operators selfadjoint relative to an indefinite metric. (Russ.), Mat. Issledovaniya, ) (4), Kishinev, Akad. Nauk MSSR., 237-40.
[8] Dissipative operators in a Hilbert space with an indefinite metric (Russ.), Izvestiaya Akad. Nauk. USSR, ser. mat., 1973, 37(3). (English translation, Math. USSR Icv., 7, 639-60.) [9] On invariant subspaces of commutative families of operators in a space with an indefinite metric (Russ.), Ukrainsk. mat. zhurnal, 1976, 278(3), 293-9. [10] On uniformly J-expansive operators and focusing plus-operators (Russ.). In the
book: Shkola po teorii operatorov v funktsional'nykh prostranstvakh, Minsk, 1978.
[11] Parametric representation of operators acting in a space with an indefinite metric (Russ.), Dep. VINITI, 1979, 1138-79.
[12] On extensions of J-isometric operators (Russ.). In the book: XVI Voronezh Summer School in Mathematics, Dep. VINITI, 1981, 5691-81, 21. [13] On the completeness and basicity of a system of eigenvectors and associated vectors of J-selfadjoint operators of the class J (H) (Russ.), Doklady Akad. Nauk, USSR, 1980, 253 (5), 1033-36. [14] On the theory of extensions of isometric and symmetric operators in spaces with an indefinite metric (Russ.), Dep, VINITI, 1982, 3420-82, 29.
[15] On a class of analytic operator-functions (Russ.). In the book: VIII Shkola po teorii operatorov v funktsional'ngkh prostranstvakh, Riga, 1983, 7-8.
Azizov, T. Ya., Gordienko, N. A., and Iokhvidov, I. S. [ 1 ] On J-unitary groups generated by J-selfadjoint operators of the classes (H) and K(H) (Russ.), Dep. VINITI, 1980, 4213-80.
Azizov, T. Ya and Dragileva, M. D. [ 1 ] On invariant subspaces of operator-commutative families of operators in a space with an indefinite metric (Russ.), Trudy ALLI. mat. Voronezh Univ., 1975, ed. 17,
3-7. Azizov, T. Ya, and Iokhvidov, E. J. [ 1] On invariant subspaces of maximal J-dissipative operators (Russ.), Mat. Zametki, 1972, 12 (6), 747-754. (English translation, Math. Notes. 12 886-9.)
Azizov, T. Ya, Iokhvidov, E. I., and Iokhvidov, I. S.
[ 1 ] On the connection between the Cayley-Neyman and the Potapov-Ginzburg transformations (Russ.). In the book: Functional Analysis, Linear Spaces, Ul'yanovsk, 1983, 3-8.
References
289
Azizov, T. Ya., and Iokhvidov, I. S.
[ 1 ] A criterion for completeness and basicity of root vectors of a completely continuous J-selfadjoint operator in a Pontryagin space H, (Russ.), Mat. Issledovaniya, 6, Kishinev, Akad. Nauk Mold, SSR, 1971.
Azizov, T. Ga, Iokhvidov, I. S., and Shtraus, V. A. [ 1 ] On normally soluble extensions of a closed symmetric operator (Russ.), Mat. Issledovaniya, 7(1), Kishinev, Akad. Nauk Mold. SSR, 1972. Azizov, T. Ya., and Kondras, T. V. [1] Criteria for the existence of maximal definite invariant subspaces for operators in a space with an indefinite metric. (Russ.). Dep. VINITI, 1977, No. 4207-77.5.
Azizov, T. Ya., and Kuznetsova. N. N. [1] On p-bases in a Pontryagin space (Russ.). Uspekhi Mat. Nauk. 1983, 38, (1).
Azizov, T. Ya., and Usvyatsova, E. B. [1] Existence of special non-negative subspace of J-dissipative operators. (Russ.), Siber. Mat. Zhurnal, 1977, 18(1). [2] On completeness and basicity of eigenvectors and associated vectors of the class K(H), Ukrainsk. Mat. Zhurnal, 1978, 30(1), 86-8. [3] On the basicity of a system of root vectors of a J-selfadjoint operator generated by a pencil L(k) = I- kG - X- 1H (Russ.), Dep. VINITI, 1980, No. 962-80, 6. Azizov, T. Ya., and Khoroshavin, S. A. [1] On invariant subspaces of operators acting in a space with an indefinite metric (Russ.), Funktion analiz i ego pril., 1980, 14(4), 1-7.
Azizov, T. Ya., and Shlyakman, M. Ya. [1] On the Kiihne-Iokhvidov-Shtraus effect (Russ.), Ukrainsk. mat zhurnal, 35 (4), 484-5. Azizov, T. Ya., and Shtraus, V. A. [1] On the completeness of a system of root vectors of a definitizable operator in a
Banach space with an Hermitian form (Russ.), Trudy N.I.I. mat., Voronezh Univ., 1972, ed. 5, 1-6. Bayasgalan, Ts.
[1] On the existence of a positive square-root of an operator positive relative to an indefinite metric (Russ.), Stadia Sci. Math. Hungary, 1980, 15, 367-79. Bennevitz, Ch. [1] Symmetric relations on a Hilbert space, Lect. Notes Math, 1972, No. 280. 212-18.
Bognar, J.
[1] On the existence of a square-root of an operator self-adjoint relative to an indefinite metric (Russ.; English summary), Magyar Tud. Acad. Kutato Int. Koze, 1961. 6(3), 351-63. [2] On an occurrence of discontinuity of a scalar product in spaces with an indefinite metric (Russ.), Uspekhi Mat. Nauk, 1962, 17, 157-9.
290
References
[3] A remark on dually strict plus-operator (Russ.), Mat. Issledovaniya, 1973, 8(1) (27), 217-19, Kishinev. [4] Non-degenerate inner-product spaces spanned by two neutral subspaces, Studia Sci. Math. Hungary, 1978, 13, 463-8. [5] A proof of the spectral theorem for J-positive operators, Acta Sci. Math., 1983, 15(1-2), 75-80.
Brodskiy, M. L.
[ 1 ] On the properties of an operator mapping into itself the non-negative part of a space with an indefinite metric (Russ.), Uspekhi Mat. Nauk, 1959, 14(1), 147-52. Davis, Ch. [1]
J-unitary dilation of general operators, Acta Sci. Math., 1971, 32, 127-99.
Davis, Ch. and Foia5, C. [1]
Operators with bounded characterstic function and their J-unitary dilation, Acta. Sci. Math., 1971, 32, 127-9.
Derguzov, V. I.
[1] On the stability of solutions of Hamiltonian equations with unbounded periodic operator coefficients (Russ.), Mat. Sbornik, 1964, 63, 591-619. [2] Necessary conditions for strong stability of Hamiltonian equations with unbounded periodic operator coefficients (Russ.), Vestnik Leningr. State Univ., 1964 (19), 18-30. [3] Sufficient conditions for stability of Hamiltonian equations with bounded operator coefficients (Russ.), Mat. Sbornik, 1964, 64(3), 419-35.
[4] Domain of stability of linear Hamiltonian equations with periodic operator coefficients (Russ.), Vestnik Leningrad. State University, 1969, 13, 20-30.
Ektov, Yu. S. J-non-negative completely continuous operators (Russ.), Trudy N.I.I. Voronezh Univ., 1975, ed. 17, 76-86. [2] Completely continuous J-non-negative operators in a J-space with two norms, Trudy, N.I.I. Mat. Voronezh Univ., 1975, ed. 20, 57-64. [1]
Etkin, A. E. [11 Generalized resolvents of a symmetric operator in a n,-space (Russ.), Funktional'nyy analiz. Ul yanovsk, 1981, ed. 16. [2] Generalized resolvents of a symmetric operator in a TI,-space (Russ.), In the book: Funktional'ny analiz. Lineynye prostranstva, Ul'yanovsk, 1983, 149-57.
Ginzburg, Yu. P.
[1] On J-non-expansive operator-functions (Russ.), Doklady Akad. Nauk USSR, 1957, 117, (2), 171-3. [2] On J-non-expansive operators in a Hilbert space (Russ.), Nauch. zap. fiz-mat. fac. Odessa Gosud. pedag. inst., 1958, 22(1).
[3] On projection in a Hilbert space with a bilinear metric, Doklad. Akad. Nauk USSR, 1961, 139(4), 775-8. [4] On subspaces of a Hilbert space with an indefinite metric, Nauch. zap. kafedr. mat., fiz. i estestvozH. Odessa Gos. ped. inst., 1961, 25(2), 3-9.
References
291
Helton, J. W. [1] Invariant subspaces of certain commuting families of operators on linear spaces
with an indefinite inner product, Doct. diss. Stanford Univ., 1968. (Dissert. Abstrs., 1969, 29(11), 42-67.) [2] Unitary operators on a space with an indefinite inner-product J. Funct. Anal., 1970, 6(3), 412-40.
[3] Operators unitary in an indefinite metric and linear-fractional transformations, Acta Sci. Math., 1971, 32(3), 261-6. Hestenes, M. R.
[1] Applications of the theory of quadratic forms in Hilbert space to the calculus of variations, Pacific J. Math., 1951, 1, 525-81.
Iokhvidov, E. I. [ 1 ] On maximal J-dissipative operator (Russ.), Sbornik studentch. nauch. rabot, ed. S (estestvennye nauki), Voronezh, 1972, 74-9. [2] On extensions of isometric operators in a Hilbert space with an indefinite metric (Russ.), Trudy N.I.I. Mat. Voronezh Univ., 1974, ed.4, 21-31.
[3] On the properties of one linear-fractional operator transformation (Russ.), Uspekhi Mat. Nauk, 1975, 30(4), 243-4. [4] On J-unitary extensions of J-isometric operators with equal and finite deficiency numbers (Russ.), Trudy N.I.I. Mat. Voronezh Univ., 1975, 17, 34-41.
[5] Description of J-unitary extensions of J-isometric operators with equal and finite deficiency numbers (Russ.), Doklady Akad. Nauk USSR, 1976, 228(4), 783-6. [6] On closures and boundedness of linear operators of certain classes (Russ.), Uspekhi Mat. Nauk, 1979, 34(5), 217-18.
[7] A criterion for the closability and boundedness of linear operators of certain classes (Russ.). In the book: Funktional'nyy analiz, Ul'yanousk, 1980, ed. 15, 90-96. [8] On linear operators which are J-non-expansive simultaneously with their adjoints. (Russ.). Dep. VINITI, 1983, No. 3284-83.
[9] On maximal J-isometric operators (Russ.). In the book: VIII shkola po teopii operatorov v funktsional'nykh prostranstbakh, tezicy dokladov, Riga, 1983, 104-05.
lokhvidov, E. I., and Iokhvidov, I. S. [1] On global inversion of a linear-fractional operator transformation (Russ.), Trudy N.I.I. Mat. Voronezh Univ., 1973, ed. II, 64-70.
Iokhdvidov, I. S. Unitary operators in a space with an indefinite metric (Russ), Zap, N.I.I. Mat. i Mekh. Khar'kov Gas. Univ. Mat. Obsch, 1949, No. 21, 79-86. On the spectra of Hermitian and unitary operators in a space with an indefinite metric (Russ.), Doklady Akad. Nauk USSR, 1950, 71(2). Unitary and selfadjoint operators in a space with an indefinite metric (Russ.), Dissertation, Odessa, 1950. Linear operators in spaces with an indefinite metric (Russ.), Trudy 3rd Vsesayuzn.
Mat. Sezda, 1958, 3, 254-61. On the boundedness of J-isometric operators (Russ.), Uspekhi Mat. Nauk, 1961, 16(4), 167-70. On some classes of operators in a space with a general indefinite metric (Russ.). In
References
292
the book: Funktional'nyy analiz i ego primenenie, Baku, Akad. Nauk AzSSR, 1961, 90-5. [7] Singular lineals in spaces with an arbitrary Hermitian-bilinear metric (Russ.), Uspekhi Mat. Nauk, 1962, 17 (4), 127-33. [8] On operators with completely continuous iterations (Russ.), Akad. Nauk USSR, 1963, 153(2), 258-61.
[9] A new characterization of II,-spaces, Tezisy nauch. konf. Odessa inzh.-stroit Instit., Odessa, 1964, p. 173 (Russ.). [10] On a lemma of K. Fan generalizing A. N. Tikhonov's principle of fixed points (Russ.), Doklady Akad. Nauk USSR, 1964, 159, 501-04. [11] On singular lineals in II,-spaces (Russ.; English summary), Ukrainsk. Mat. Zhurnal, 1964, 16(3), 300-08. [12] G-isometric and J-semi-unitary operators in a Hilbert space, (Russ.), Uspekhi Mat. Nauk, 1965, 20 (3), 175-81. [13] On maximal definite lineals in a Hilbert space with a G-metric (Russ.), Ukrainsk. Mat. Zhurnal, 1965, 17 (4). [14] Linear-fractional transformations of J-non-expansive operators (Russ.), Doklady Akad. Nauk, Arm. SSR, 1966, 42(1), 3-8.
[15] Unitary extensions of isometric operators in the Pontryagin space II, and extensions in the class .0, of finite sequences of the class e0,,,, (Russ.), Doklady Akad. Nauk USSR, 1967, 173(4). [16] On the spectral trajectories generated by unitary extensions of isometric operators of a translation in the finite-dimensional space IT, (Russ.), Doklady Akad. Nauk USSR, 1967, 173(5), 1002-5. [17] On Banach spaces with a J-metric and certain classes of linear operators in these spaces (Russ.), Izvestija Akad. Nauk MSSR, series fiz. tekh u met. nauk, 1968,
No. 1, 60-80. [18] On a class of linear-fractional operator transformations (Russ.). In the book: Sbornik statey po funktsional'nym prostranstvam i operatornym uravnenigam, Voronezh, 1970, 18-44. [19] On spaces of finite sequences with a non-degenerate Hermitian form (Russ.), Trudy N.I.I. Mat. Voronezh Univ., 1975, ed. 20, 30-8. [20] On certain new ideas and results in the theory of linear operators in IIy-spaces (Russ.). In the book: Teoriya operatornykh urovneniy, Voronezh: Izdat. Veronezh Univ., 1979, 33-44. [21] Regular and projectionally complete lineals in spaces with a general Hermitianbilinear metric (Russ.), Doklady Akad. Nauk USSR, 1961, 139(4), 791-4.
Iokhvidov, I. S., and Ektov, Yu. S. [1]
Integral J-non-negative operators and weighted integral equations (Russ.),
Doklady Akad. Nauk Ukr. SSR. 1981. No. 6, 15-19.
[2] Equivalence of three definitions of Hermitian-non-negative kernels and a functional realization of the Hilbert spaces generated by them (Russ.), Dep. VINITI, 1983, No. 2834-83.
Jonas, P. [ 1 ] A condition for the existence of an eigen-spectral function for certain automorphisms of locally convex spaces (Ger.), Math. Nachr., 1970, 45, 143-60. [2] On the maintenance of the stability of J-positive operators under J-positive and negative perturbations (Ger.), Math. Nachr, 1975, 65. [3] On the existence of eigen-spectral functions for J-positive operators, I (Ger.), Math. Nachr., 1978, 82, 241-54.
References
293
[4] On the functional calculus and the spectral function for definitizable operators in Krein space, Beltrage Anal., 1981, 16, 121-35. [5] Compact perturbations of definitiable operators, II, J. Operator Theory, 1882, 8, 3-18. [6) On spectral functions of definitizable operators, Banach Centre Publications, 1982, 8, 301-11. [7] On a class of J-unitary operators in Krein space, Preprint Akad. der Wiss. der DDR, Inst. fur Mathematik, Berlin, 1983.
Jonas, P. and Langer, H. [1] Compact perturbations of definitizable operators, J. Operator Theory, 1979, 2, 63-77. [2] Some questions on the perturbation theory of J-non-negative operators in Krein spaces, Math. Nachr., 1983, 114, 205-26. Khastskevich, V. A.
[11 On some geometrical and topological properties of normed spaces with an indefinite metric (Russ.), Izvestiya Vuzov, Matematika, 1973.
[2] Remarks on dually plus-operators in a Hilbert J-space (Russ.). In the book: Funktsional'nyy analiz i ego prilozheniya, ed. I, Voronezh, 1973, 27-37. [3) On plus-operators in a Hilbert J-space (Russ.), Mat. Issledovaniya, 9, (2(32)), Krishinev, Akad. Nauk MSSR, 1974, 182-203. [4] On J-semi-unitary operators in a Hilbert J-space (Russ.), Mat. Zametki, 1975, 17(4), 639-47. [5] On fixed points of generalized linear-fractional transformations (Russ.), Izvestiya Akad. Nauk, USSR, set. Mat., 1975, 39(5), 1130-41. [6] On an application of the principle of compact mappings in the theory of operators in a space with an indefinite metric (Russ.), Funktsion. analiz i ego pril., 1978, 12(1), 88-9. [7) Focusing and strictly plus-operators in Hilbert space (Russ.), Dep. VINITI, 1980, No. 1343-80.
[8] On the symmetry of properties of a plus-operator and its adjoint (Russ.), Funktsion. analiz, ed. 14, Ul'yanovsk, 1980, 177-186. [9] On strict plus-operators (Russ.), Dep. VINITI, 1980, No. 2440-80. [10] Uniformly definite invariant dual pairs of plus-operators with a quasi-focusing power (Russ.), Dep. VINITI, 1981, No. 1917-81. [11] On invariant subspaces of focusing plus-operators (Russ.), Mat. Zametki, 1981, 30(5), 695-702.
[12] Invariant subspaces of plus-operators with a focusing power (Russ.), Dep. VINITI, XV Voronezh winter Mat. School. [13] A generalized Poincare metric on an operator ball (Russ.), Funktsion. analiz i ego pril., 1983, 17 (4), 93-94. [14] On a generalization of the Poincare metric for a unit operator ball (Russ.), Dep. VINITI, 1983, No. 6913-83.
[15] Invariant subspaces and properties of the spectrum of plus-operators with a quasi-focusing power (Russ.), Funktsian analiz i ego pril., 1984, 18(1).
Kholevo, A. S.
[1] A generalization of Neyman's theory about the operator T"T on spaces with an indefinite metric (Russ.), Uchen. zapis Azerb. Univ. Ser. fiz-mat. nauk, 1965, No. 2, 45-8.
References
294 Kopachevskiy, N. D.
[1] Theory of small osicallations of a liquid taking surface tension and rotation into account, Dissertation, Khar'kov, 1980 (Russ.). [2] On the properties of basicity of a system of eigenvectors and associated vectors of
a selfadjoint operator bundle I- kA - k-'B, (Russ.), Funktsion. analiz i ego pril., 1981, 15(2).
[3] On the p-basicity of a system of root vectors of a selfadjoint operator bundle I - XA - X 'B (Russ.). In the book: Funktsion. analiz iprikladnaya matematika, Kiev, 1982, 55-70.
Kostguchenko, A. G. and Orazov, M. B.
[1] On some properties of the roots of a selfadjoint quadratic bundle (Russ.), Funktsion. analiz i ego pril., 1975, 9(4), 28-40.
Krasnosel'skiy, M. A. and Sobolev. A. V. [1] On cones of finite rank (Russ.), Doklady Akad. Nauk USSR, 1975, 225(6).
Krien, M. G.
[1] . On weighted integral equations the distribution functions of which are not monotonic (Russ.), In the memorial volume to D. A. Grave, Moscow, 1940, 88-103. [2] On linear completely continuous operators in functional spaces with two norms (Ukr.), Zhurnal Akad. Nauk Ukr. RSR, 1947, 9, 104-29. [3] Helical curves in an infinite-dimensional Lobachevskiy space and Lorents transformation (Russ.), Uspekhi Mat. Nauk, 1948, 3 (3). (Amer. Math. Soc. Transl. (2), 1, 1955, 27-35.)
[4] On an application of the fixed-point principle
in
the theory of linear
transformations with an indefinite metric (Russ.), Uspekhi Mat. Nauk. 1950. 5(2) 180-90.
[5] A new application'of the fixed-point principle in the theory of linear operators in a space with an indefinite metric (Russ.), Doklady Akad. Nauk. USSR, 1964, 154(5), 1026-26. English translation: Soviet Math. Doklady, 5 (1964), 224-7. [6] On the theory of weighted integral equations (Russ.), Izvestiya Akad. Nauk. MSSR, 1965 (7), 40-6.
Krein, M. G. and Langer, G. K. [ 1 ] On the spectral function of a self-adjoint operator in a space with an indefinite
metric (Russ.), Doklady Akad. Nauk USSR,
1963, 152,
39-42. English
translation: Soviet Math. Doklady 4, 1963, 1236-9. [2] On some mathematical principles of the linear theory of damped oscillations of continua (Russ.). In the book: Trudy mezhdunarod. simposiuma po prilozheniyam meopii funktsiy u mechanike splashnoy sredy, vol. 2. Moscow, Nauka, 1966, 283-322. [3] On definite subspaces and gneralized resolvents of an Hermitian operator in a II,-space (Russ.), Funktsion. analiz i ego pril., 1971, 5 (2), 59-71; 1971, 5(3), 54-69. [4] On the generalized resolvents and the characteristic function of an isometric operator in IL-spaces (Ger.), Colloquia Math. Soc. Janos Bolyai, vol. 5; Hilbert space operators and operator algebras, Amsterdam, London, North-Holland, 1972, 353-99.
References
295
[5] On the Q-function of a a-hermitian operator in II,-spaces (Ger.), Acta Sci. Math., 1973, 34, 191-230.
[6] On some extension problems closely connected with the theory of Hermitian operators in IIh-spaces, I: Some classes of functions and their representations (Ger.), Math. Nachr., 1977, 77, 187-236; II. Generalized resolvents, u-resolvents and entire operators (Ger.), J. Funct. Analysis, 1978, 30, 3, 290-447.
[7] On some extension problems closely connected with the theory of Hermitian operators in a II,-space, III: Indefinite analogue of the Hamburger and Strieltjes moment problems, Beitrage zur Analysis, 1979, 14, 25-40; 1981, 15, 27-45.
Krein, M. G. and Rutman, M. A. [1]
Linear operators leaving invariant a cone in a Banach space (Russ.), Uspekhi Mat.
Nauk, 1948, 3 (1), 3-93; Amer. Math. Soc. Transl. 1950, 26.
Krein, M. G. and Shmul'yan, Yu. L. [11 On a class of operators in a space with an indifinite metric (Russ.), Doklady Akad. Nauk USSR, 1966, 170(1), 34-37; Soviet Math. Dokl. 1966, 7, 1137-1141. [2] On plus-operators in a space with an indefinite metric (Russ.), Mat. Issledovaniya, 1966, 1 (1), 131-61 Kishinev; Amer. Math. Soc. Transl., 2(8), 1969, 93-113.
[3] J-polar representation of plus-operators (Russ.), Mat. Issledovan, 1966. 1 (2), 172-210, Kishinev; Amer. Math. Soc. Transl. 2(85), 1969, 115-43 [4] On linear-fractional transformations with operator coefficients (Russ.), Mat. Issledovaniya, 1967, 2(3) 64-96, Kishinev.
[5] On stable plus-operators in J-spaces (Russ.). In the book: Lineynye operatory, Kishinev, Shtiintsa, 1980, 67-83.
Krein, S. G. [1] Oscillations of a viscous fluid in a container (Russ.), Doklady Akad. Nauk, USSR, 1964, 159(2), 262-5.
Kiihne, R. [1] On a class of J-self-adjoint operators (Ger.), Math. Annal., (1964), 154(1), 56-69.
Kuzhel, A. V.
[1] Spectral analysis of quasi-unitary operators of first rank in a space with an indefinite metric (Ukr.), Dopovidi Akad. Nauk Ukr. RSR, 1962, 5, 572-4. [2] Spectral decomposition of quasi-unitary operators of arbitrary rank in a space with an indefinite metric (Ukr.), DopovidiAkad. Nauk Ukr. RSR, 1963,4,430-3. (Russ. and English summaries)
[3] Spectral analysis of bounded non-selfadjoint operators in a space with an indefinite metric (Russ.), Doklady Akad. Nauk USSR, 1963, 151, 772-4. [4] On one case of the existence of invariant subspaces of quasi-unitary operators in a space with an indefinite metric (Ukr.), Dopovidi Akad. Nauk Ukr RSR, 1966, 5, 583-585. [5] Spectral analysis of quasi-unitary operators in a space with an indefinite metric (Russ.). In the book: Teoriga funktsiy, funktsronal'nyy analiz i ikh prilozheniya, 4, Khar'kov, 1967.
[6] Spectral analysis of unbounded non-selfadjoint operators in a space with an indefinite metric (Russ.), Doklady Akad. Nauk USSR, 1968, 178, 31-3. [7] Regular extensions of Hermitian operators in a space with an indefinite metric (Russ.), Doklady Akad. Nauk USSR, 1982, 265 (5).
296
References
[8] Selfadjoint and J-selfadjoint dilatations of linear operators (Russ.). In the book: Teoriya funktsiy, funktsional'nyy analiz i ikh pril., 37, Khar'kov, 1982, 54-62. [9] J-selfadjoint and J-unitary dilatations of linear operators (Russ.), Funktsion. analiz i ego pril., 1983, 17 (1), 75-6.
Langer, H. On J-Hermitian operators (Russ.), Doklady Akad. Nauk USSR, 134(2), 263-6, 1960.
On the spectral theory of J-selfadjoint operators (Ger.), Math. Ann. 1962, 146(1), 60-85.
A generalization of a theorem of L. S. Pontryagin (Ger.), Math. Ann. 1963, 152(5), 434-6. Spectral theory of linear operators in J-spaces and some applications to the bundle L(X) = X21+ XB + C, Habilitationschrift, Dresden, 1965 (Ger.). Spectral functions of a class of J-selfadjoint operators (Ger.), Math. Nachr. 1967,
33(1-2), 107-20. On strongly damped bundles in Hilbert space (Ger.), J. Math. Mech, 1968, 17, N7, 685-705. On maximal dual pairs of invariant subspaces of J-selfadjoint operators (Russ.), Mat. Zametki, 1970, 7 (4), 443-7. Generalized resolvents of a J-non-negative operator with finite deficiency (Ger.), J. Funct. Analysis, 1971, 8, 287-320. Invariant subspaces of definitizable J-selfadjoint operators (Ger.), Ann. Acad. Sci. Fenn., 1971, Ser. Al, No. 475. On a class of polynomial bundles of self-adjoint operators in Hilbert space (Ger.), J. Funct. Analysis, 1973, 12(1), 13-29; 1974, 16, 921-34. Invariant subspaces for a class of operators in spaces with indefinite metric, J. Funct. Analysis, 1975, 19, (3), 232-41. On the spectral theory of polynomial bundles of self-adjoint operators, (Ger.), Math. Nachr. 1975. 65, 301-19. On invariant subspaces of linear operators acting in a space with an indefinite metric (Russ.), Doklady Akad Nauk, USSR, 1966, 169, 12-15.
Langer, H. and Nayman, B. [1] Perturbation theory for definitizable operators in Krein spaces. J. Operator Theory, 1983, 9, 297-317.
Langer, H. and Sorjonen, P. [1]
Generalized resolvents of Hermitian and isometric operators in a Pontryagin sapce, Ann. Acad. Sci. Fenn., 1974, Ser. Al. No. 561 (Ger.).
Larionov, V. A. [1] Extensions of dual subspaces (Russ.), Doklady Akad. Nauk, USSR, 1967, 176(3), 515-17. [2] On a commutative family of operators in a space with an indefinite metric (Russ.), Mar. Zametki, 1967, 1(5), 589-94.
[3] Extensions of dual subspaces invariant relative to an algebra (Russ.), Mat. Zametki, 1968, 3(3), 253-60. [4] On the number of fixed points of a linear-fractional transformation of an operator ball on to itself (Russ.), Mat. Sbornik, 1969, 78(2), 202-13.
[5] On self-adjoint quadratic bundles (Russ.), Izvestiya Akad. Nauk, USSR, Ser. Matem., 1969, 33(1), 138-54.
References
297
[6] On bases composed of eigenvectors of an operator bundle (Russ.), Doklady Akad. Nauk USSR, 1972, 206(2), 283-86. [7] Localization of the spectrum and dual completeness of the normal motions in the problem of motion of a viscous liquid subject to surface-tension forces (Russ.), Doklady Akad. Nauk USSR, 1974, 217(3), 522-25. [8] On the principles of localization of frequencies and the principle of completeness and dual completeness of the root functions in problems of the linear theory of vibrations in continuous media (Russ.), Trudy TsNII building construction, 1974,
4, 11-20. [9] On the theory of linear operators acting in a space with an indefinite metric (Russ.). In the book: Trudy letney shkoly po spectral'noy teorii operatorov i predstavleniy grupp, 1968. Baku: Elm, 1975, 140-48.
Louhivaara, I. S. [ 1 ] Remark on the theory of Nevanlinna spaces (Ger.), Ann. Acad. Sci. Fenn., 1958, Ser. Al, No. 232. [2] On the theory of subspaces in linear spaces with indefinite metric, (Ger.), Ann. Acad. Sci. Fenn., 1958, Ser. Al, No. 252. [3] On various metrics in linear spaces (Ger.), Ann. Acad. Sci. Fenn, 1960, Ser. AI, No. 282.
Markin, S. P. [1] On the separability of linear spaces with a sesquilinear form (Russ.), Trudy Mat. Fak. Voronezh Univ., 1973, 10, 107-12.
Masuda, K.
[1] On the existence of invariant subspaces in spaces with indefinite metric. Proc. Amer. Math. Soc., 1972, 32(2) 440-4. McEnnis, B. W.
[1] Fundamental reducibility of selfadjoint operators on Krein space, J. Operator Theory, 1982, 8, (2), 219-25. Naymark, M. A. [1] On commutative unitary operators in rIb-space (Russ.), Doklady Acad. Nauk USSR, 1963, 149(6), 1261-3. [2] On commuting unitary operators in a space with an indefinite metric Acta Sci. Math., 1963, 24(3-4), 177-89. [3] An analogue of Stone's theorem in a space with an indefinite metric (Russ.), Doklady Nkad. Nauk USSR, 1966, 170(6), 1259-61; Soviet Math. Doklady, 1966, 7, 1366-8.
Nevanlinna, R. Extensions of the theory of Hilbert space (Ger.), Comm. Sen. Math. Univ. Lund. tome supplementaire, 1952, 160-8. [2] On metric linear spaces, II: bilinear forms and continuity (Ger.), Ann. Acad. Sci. Fenn., 1952, Ser. Al, No. 113. [3] On metric linear spaces, III; theory of orthogonal systems (Ger.), Ann. Acad. Sci. Fenn., 1952, Ser. Al, No. 115. [4] On metric linear spaces, IV: on the theory of subspaces (Ger.), Ann. Acad. Sci. Fenn., 1954, Ser. AI, No. 163. [1]
298
References
[5] On metric linear spaces, V: relations between different metrics (Ger.), Ann. Acad. Sci. Fenn., 1956, Ser. Al, No 22.
Noel, G. [1] Strongly J-normal operators in a J-space (Fr.), Bull. Sci. Acad. Roy. Belg., 1965, 51(5), 570-85. Ovchinnikov, V. I. [ 1 ] On the decomposition of spaces with indefinite metric (Russ.), Mat. Issled., 1968, 3, 4, 175-7, Kishinev. Pesonen, E.
[1] On the spectral representation of quadratic forms in linear spaces with indefinite metric (Ger.), Ann. Acad. Sci. Fenn., 1956, Ser. Al, No. 227. Phillips R.
Dissipative operators and hyperbolic systems of partial differential equations, Trans. Amer. Math. Soc., 1959, 90(2), 193-254. [2] Dissipative operators and parabolic partial differential equations, Communs. Pure and Applied Math., 1959, 12(2), 249-76. [3] The extension of dual subspaces invariant under an algebra, Proc. Internat. Sympos. Linear Spares, Jerusalem, 1960. Pergamon Press, 1961, pp. 366-98. [1]
Pontryagin, L. S. [1]
Hermitian operator in spaces with indefinite metric (Russ.), Izvestiya Akad. Nauk USSR, Ser. Matem, 1944, 8, 243-80.
Potapov, V. D.
[ 1 ] The multiplicative structure of J-contractive matrix-functions (Russ.). Trudy Moskov. Mat. Obshch., 1955, 4, 125-236; Amer. Math. Soc. Transl. (2), 15, 1960, 131-243.
Prigorskiy, V. A.
[ 1 ] On some classes of bases of a Hilbert space (Russ.), Uspekhi Mat. Nauk, 1965, 20(5), 231-6. Ritsner, V. S.
[1] Extension of J-Hermitian operators in J-spaces (Russ.). In the book: Issledovanie
roboty kleenykh derevyannykh konstruktsiy. Khabarovsk. Kh.
P.
I., 1975,
186-95. [2] Generalized resolvents of a non-densely defined accretive operator in a Hilbert space with indefinite metric, (Russ.). In the book: Materialy nauchotekhnicheskoy konferentsii molodykh uchenykh, Khabarovsk, 1977, 128-9.
[3] Extension of dual pairs of subspaces (Russ.), Funktsion, analiz, Ul yanovsk, 1980, 15, 135-9. [4] Theory of linear relations (Russ.), Dep. VINITI, 1982, No. 846-82.
Sakhnovich, L. A. [1] On a J-unitary dilatation of a bounded operator (Russ.), Funkision. analiz i ego pril., 1974, 8(8), 83-4.
References
299
Savage, L. J.
[ 1 ] The application of vectorial methods to metric geometry, Duke Math. J., 1946, 13, 521-28.
Scheibe, E.
[1] On Hermitian forms in topological vector-spaces, (Ger.), I. Ann. Acad. Sci. Fenn., 1960, Ser. Al, No. 294. Shmul'yan, Yu. L. [1] On division in the class of J-non-expansive operators, (Russ.), Mat. Sbornik, 1967, 74, 516-25. [2] On linear-fractional transformations with operator coefficients and operator balls (Russ.), Mat. Sbornik, 1968, 77 (3).
[3] On J-non-expansive operators operators in J-spaces (Russ.), Ukrainsk. Mat. Zhurn., 1968, 20 (3), 352-62. [4] Theory of extensions of operators and space with an indefinite metric (Russ.), Izvestiya Akad. Nauk USSR, Ser. Mat., 1974, 38(4). [5] Theory of linear relations and space with an indefinite metric (Russ.), Funktsion. analiz i ego pril., 1976, 10(1).
Generalized linear fractional transformations of operator balls (Russ.), Siber. Mat. Zhurn, 1978, 19(2) 418-25. [7] On transformers of linear relations in J-spaces (Russ.), Funktsion. analiz i ego pril., 1980, 14(2) 39-44. [8] Generalized linear-fractional tranformations of operators balls (Russ.), Siber. Mat. Zhurn., 1980, 21(5), 114-31. [6]
Shtraus, A. V. [1]
Extensions and generalized resoluents of a non-densely defined symmetric operator (Russ.), Izvestiya Akad. Nauk USSR, Ser. Mat., 1970, 34(1).
Shtrauss, V. A. [1] The Gram operator and G-orthonormalized systems in a Hilbert space, Sbornik trad. aspirantov, Mat. Fac. Voronezh Univ., 1971, 1, 85-90. [2] G-orthonormalized systems and bases in a Hilbert space (Russ.), Izvestiya Vuzov. Matematika, 1973, No. 9. 108-17. [3] On the theory of selfadjoint operators in Banach spaces with an Hermitian form (Russ.), Siber. Mat. Zhurnal, 1978, 19(3). [4] On the integral representation of a J-selfadjoint operator in II, (Russ.), Sbornik nauch. trudov No. 252: prikladnaya matematika. Chelyabinsk, 1980, 114-18.
Shtraus, V. A. and Ektov, Yu. S. [1] An analogue of Dini's theorem and J-non-negative operators (Russ.), Funktsion. analiz Ul'yanovsk, 1980, 15, 188-91.
Shul'man, V. S.
[1] On fixed points of linear-fractional transfomations (Russ.), Funktsion. analiz i ego pril., 1980, 14 (2), 93-4.
References
300
Sobolev, A. V. and Khatskevich, V. A.
[1] Invariant subspaces and properties of the spectrum of a focusing plus-operator (Russ.), Rukopis dep. VINITI, 1980, No. 1344-80. [2] On definite invariant subspaces and the structure of the spectrum of a focusing plus-operator (Russ.), Funktsion. analiz i ego pril., 1981, 15(1), 84-5. Sobolev, S. L.
[ I] Motion of a symmetric top with a cavity filled with liquid (Russ.), ZhPMTF. 1960,
3, 20-55. Spitkovskiy, I. M.
[ 1 ] On the block structure of J-unitary operators (Russ.). In the book: Teoriya funkisiy, funktsional'nyy analiz i ikh pritozheniya, Kharkov, 1978, 129-38.
Usvyatsova, E. B.
[1] On the spectral theory of G-isometric operators (Russ.), Mat. Fak. Voronezh Univ., 1972, 66-9. [2] On the structure of root subspaces of J-dissipative operators (Russ.). In the book: Melody resheniya operatornykh uravneniy, Voronezh, 1978, 158-60. Wittstock, G. [ 1 ] On invariant subspaces of positive transformations in spaces with indefinite metric
(Ger.), Math. Ann, 172(3), 167-75.
INDEX
Adjoint operator 49 Almost J-orthonormalized basis 76 Almost normalized system of vectors 76 Amenable group of operators 137 Angular operator 49 Anti-linearity 2 Anti-space 4 Aperture of two subspaces 61 Asymptotically isotropic sequence 42 B-plus-operator 132
Biorthogonal systems of vectors 76 Bounded spectral set 97
Canonical decomposition of a space 14 Canonical imbedding 183 Canonical imbedding operators 22 Canonical projectors 18 Canonical symmetry operator 21 Carrier a(E) of homomorphism 210 Cauchy-Bunyakovski inequality 6 Cayley-Neyman transformation 142 Centralized system of subsets 59 Class H operators 193 Class K(H) operators 197 Closed mapping in a Hausdorff space 177
Complete system of vectors 219 Completely invariant subspace relative to a family of operators 161 Condition L 160 Condition A+ 170 Conditional basis 77 Continuous spectrum of a linear relation 145
`Corner' 170 Critical points 220
Decomposable lineal 34 Deficiency subspace 25 Definite lineal 3 Definite subspace 24 `Definitizable' J-selfadjoint operator 211 Degenerate lineal 5 Dilatation of an operator 266 Direct Cayley-Neyman transformation 144
Dissipative kernel 234 Dissipative operators 91
Domain of definition of a linear relation 85
Doubly strict plus-operator 120 Dual pair 72
Eigen spectral function of a J-selfadjoint operator 211 Eigenlineal of a linear relation 143 Eigenvalues of a linear relation 143 Eigenvector of L(X) 237 Eigenvectors of a linear relation T 143 Extension of a dual pair 72 Field of regularity of a linear relation 44 Field of regularity of an operator 92 Fixed point of the transformation F; 176
Focusing plus-operator 126 Function generated by an operator 268 Function T(µ) belonging to the class ,r(. j,IY2) 268
G-adjoint of a linear relation 146 G-metric 40 G-selfadjoint linear relation 146 G-selfadjoint operator 104 301
302
Index
G-space 40 G (')-space 69 G-symmetric linear relation 104 G-symmetry 101 (G1, G2)-adjoint 85 (G,, G2)-semi-unitary operator 35
(G,, G2)-unitary operator 135 Generalized linear-fractional transformation 176 Generalized resolvent of a J-symmetric operator 279 Gram operator 39 Hermitian form 1 Hermitian kernel 269 Hermitian non-negative kernel 234 Hermitian operator 104 Hermitian positive kernel 234 Hermitian symmetric 2 Hilbert-Schmidt operator 222 Hyper-maximal neutral lineal 28
J-non-contractive operator 117 J-non-negative 107 J-orthogonal complement 21 J-orthogonal projection 45 J-orthogonality 21 J-orthonormalized bases 72 J-orthonormalized system 74 J-positive 107 J-real part of an operator 96
J-space 20
J-spectral function with a set s(E) of critical points 210 J-symmetry 20 Ji-module 216
(Ji, Jz)-bi-extension of an operator 248 (J1, Jz)-bi-non-contractive operator 124 (J,, Jz)-isometric operator 140 (Ji, Jz)-non-expansive operator 129 (Ji, Jz)-semi-unitary operator 134 (Jl, Jz)-unitary dilatation of an operator 266
(J1, Jz)-unitary operator 141 Identical linear relation 143 Indecomposable lineal 34 Indefinite form 2 Indefinite lineal 3 Indefinite metric 2 Indefiniteness 85 Indefiniteness of a linear relation 143
Jordain chain of associated vectors 237
Kernel 85
Kernel of a linear relation 85 Krein space 14 Krein-Shmul'yan linear-fractional transformation 176
Index of a subspace T, 256 Index of a (D-operator 94 Inertia index 38 Intrinsic completeness 32 Intrinsic metric 31 Intrinsic norm 31 Invariant subset relative to a generalized linear-fractional transformation 176 Invariant subspace 158 Inverse Cayley-Neyman transformation
(t+, 'P_ )-decomposition of a subspace 73
Law of inertia 38 Lineal 2 Linear 1
Linear dimension 24 Linear relation 25 Linear space 1 Linearly ordered subset 7
144
U (.Y°) 51
Involution 19
Isometrical isomorphism 4 Isotropic lineal 5 Isotropic vector 5
J-accumulative linear operator 147 J-bi-extension of a J-Hermitian operator 264 J-complete J-orthonormalized system 74 J-accumulative linear operator 149 J-form 21 J-imaginary part of an operator 96 J-isometricity 20 J-metric 20
Maximal completely invariant subspace 166
Maximal dual pair 72 Maximal extension of a pair 72 Maximal invariant dual pair 166 Maximal invariant subspace 166 Maximal J-accumulative operator 147 Maximal J-orthonormalized system 74 Maximal negative lineal 6 Maximal neutral lineal 6 Maximal neutral subspace 28 Maximal non-degenerate lineal 6 Maximal positive lineal 6
Index
303
Maximal W-accumulative operator 147 Maximal W-dissipative operator 91 Maximal W-symmetric operator 104 Minus-operators 129
Q-orthogonal sets 5 Q-orthogonal vectors 5 (Q1, Q2)-isometric isomorphism 4 (Q,, Q2)-isometrically isomorphic spaces
Negative lineal 3 Negative vector 3 Neutral lineal 3 Neutral vector 3 Non-closed definite lineal 32 Non-decreasing linear operators 155 Non-degenerate lineal 5 Non-negative Hermitian kernel 242 Non-negative lineal 3 Non-negative vector 3 Non-positive lineal 3 Non-positive vector 3 Non-strict plus-operator 118 Normal eigenvalue 89 Normal point 89 Normalized system of vectors 76 Normally decomposable operator 139 Nuclear operator 222
(Qt, Q2)-skew-symmetric isomorphism 4 Quadratic bundle 238 'Quasi-inverse' operator 107
4
Orthogonal canonical orthoprojectors 21 Orthogonal sum Q+ 15 Orthonormalized basis 219 Ortho-projectors 18 P-basis 226 PG-transformation 246 Partial ordering by inclusion 7 Partially (J,, J2)-isometric operators 140 Plus-deficiency 122 Plus operator 117 Point of definite (positive or negative) type 210 Point of regular type 92 Point of spectrum of L(X) 237 Point spectrum vp 143 Point spectrum of a linear relation 143 Pontryagin space H, 64 Positive lineal 3 Positive vector 3 Potapov-Ginzburg transformation 245 Projectionally complete lineal 44 Property (D 165 Property 165 Property 1'J 165 Q-biothogonality 10 Q-metric 7 Q-orthogonal complement 5 Q-orthogonal direct sum 14
Range of indefiniteness 38 Range of values of a linear relation 85 Rank of indefiniteness of a subspace 38
Regular critical point of a J-spectral function 211 Regular definite lineal 34 Regular definite subspace 30 Regular dilatation of an operator 266 Regular extension 262 Regular G-metric 40 Regular generalized resolvent 279 Regular J-spectral function 211 Regular points of a linear relation 143 Regular point of L(X) 237 Regular subspace 46 Residual spectrum of a linear relation 144
Resolution of the identity 36 Resolvent set for a linear relation 143 Riesz basis 77 Riesz projector 97 Root lineal 88 Root vector 88
S-bounded operator 94 S-completely continuous operator 94 S-continuous operator 94 s-number of an operator 222 sp A 223 Y, 222 Scalarly commutative family 206 Selfadjoint operator 16 Semi-bounded below 107 Semi-definite lineal 24 Semi-definite spaces 24 Semi-definite subspace 24 Semi-linearity 2 Sesquilinear form 1 Set of critical points of an operator 211 Simple J-dissipative operator 224 Simple J-unitary operator 278 Single-valued linear relation 143 Singular critical point 211 Singular definite lineal 34 Singular G-metric 40
Index
304
Skew-symmetric isomorphism 4 Skewly linked lineals 10 Spectral function 36 Spectrum a
143
Stable J-unitary operator 137 Stable plus-operator 131 Standard operator 253 Strict plus-operator 118 Strictly dissipative kernel 234 Strictly J-dissipative operator 102 Strongly continuous on the right 36 Strongly damped bundle 241 Strongly stable J-unitary operator 139 Subspaces: dual pair 72 Subspaces: extension of a dual pair 72 Subspaces: maximal dual pair 72 Subspaces: maximal extension of a dual pair 72 Subspaces of the h` class 33 Sum of two linear relations 143 Symmetric operator 104 Trace of an operator 223 Unconditional basis 77 Uniformly definite lineal 30 Uniformly J-dissipative operator 102 Uniformly J-positive 102 Uniformly (J1, J2)-bi-expansive operator 126
Vectors associated with an eigenvector of L(X) 237 W-accumulative linear operator 147 W-accumulative linear relation 147 W-dissipative linear relation 147 W-dissipative operator 91 W-Hermitian operator 104 W-isometric linear relation 146 W-metric 39 W-non-contractive linear relation 146 W-non-expansive linear relation 146 W-space 39
W-symmetric linear relation 147 W-symmetric operator 104 W(')-space 69 (W,, W2)-isometric operator 134 (W1, W2)-non-contractive operator 117 (W,, W2)-semi-unitary operator 134 (W,, W2)-unitary operator 134
x-regular dilatation of an operator 266 x-regular extension 262 x-regular generalized resolvent 279 x-regular Xo-standard extension 264 Xo-standard set of closed J-Hermitian operators 264 a-orthogonal complement 67 a-non-negative kernel 242
Uniformly (J1, J2)-expansive operator 126
Uniformly negative lineal 31 Uniformly positive lineal 31
(D-operator 94 4 o-operator 94 (Do-point of an operator 94