Author: Kothe G.  

Tags: mathematics   vectors  

Year: 1969

Text
                    Die Grundlehren der
mathematisdien Wissenschaften in Einzeldarstellungen
Band 159
Gottfried Kothe
Topological Vector Spaces I


Die Grundlehren der mathematischen Wissenschaften in Einzeldarstellungen mit besonderer Beriicksichtigung der Anwendungsgebiete Band 159 Herausgegeben von J. L. Doob • A. Grothendieck • E. Heinz • F. Hirzebruch E. Hopf • H. Hopf • W. Maak • S. MacLane • W. Magnus M. M. Postnikov • F. K. Schmidt • D. S. Scott • K. Stein Geschaftsfuhrende Herausgeber B. Eckmann und B. L. van der Waerden
Gottfried Kothe Topological Vector Spaces I Translated by D. J. H. Garling i Springer-Verlag New York Inc. 1969
Prof. Dr. Dr. h.c. Gottfried Kothe Institut fur angewandte Mathematik der Johann-Wolfgang-Goethe-Universitat, Frankfurt am Main Geschaftsfiihrende Herausgeber: Prof. Dr. B. Eckmann Eidgenossische Technische Hochschule Zurich Prof. Dr. B. L. van der Waerden Mathematisches Institut der Universitat Zurich Translation of Topologische Lineare Raume I, 1966 (Grundlehren der mathematischen Wissenschaften, Vol. 107) All rights reserved. No part of this book may be translated or reproduced in any form without written permission from Springer-Verlag. © by Springer-Verlag Berlin • Heidelberg 1969. Library of Congress Catalog Card Number 78-84831 Printed in Germany Title No. 5142
Preface to the First Edition It is the author's aim to give a systematic account of the most important ideas, methods and results of the theory of topological vector spaces. After a rapid development during the last 15 years, this theory has now achieved a form which makes such an account seem both possible and desirable. This present first volume begins with the fundamental ideas of general topology. These are of crucial importance for the theory that follows, and so it seems necessary to give a concise account, giving complete proofs. This also has the advantage that the only preliminary knowledge required for reading this book is of classical analysis and set theory. In the second chapter, infinite dimensional linear algebra is considered in comparative detail. As a result, the concept of dual pair and linear topologies on vector spaces over arbitrary fields are introduced in a natural way. It appears to the author to be of interest to follow the theory of these linearly topologised spaces quite far, since this theory can be developed in a way which closely resembles the theory of locally convex spaces. It should however be stressed that this part of chapter two is not needed for the comprehension of the later chapters. Chapter three is concerned with real and complex topological vector spaces. The classical results of Banach's theory are given here, as are fundamental results about convex sets in infinite dimensional spaces. The subsequent chapters contain a full account of the properties of locally convex spaces. This account is concerned above all with the general theory, but some important classes of spaces, such as for example (F)-spaces, barrelled spaces and bornological spaces, are considered in greater detail. A large number of examples and counterexamples are intended to enable both the scope and the limits of the theory to be seen. The second volume will contain the theory of linear mappings and the special spaces and classes of spaces which are important in analysis. The theory of Hilbert space will not be dealt with, since there are plenty of excellent textbooks on this topic. Information about the contents of the book is given in the detailed table of contents at the beginning of the book, and in the short summaries at the beginning of each chapter. No claim for completeness is made
VI Preface to the First Edition for the bibliography at the end of the book, but it should nevertheless be detailed enough to enable further independent work to be done. My teacher O. Toeplitz provided the first impulse for work on the theme of this book. In § 30, I have endeavoured to give an account of the theory of perfect spaces, which was developed by us together, I have to thank repeated personal contact with my French colleagues J. Dieudonne, A. Grothendieck and L. Schwartz since the war, for detailed knowledge of the theory developed by them; this forms the main subject-matter of this book. The present account is frequently based on the two volumes of Bourbaki (Bourbaki [6] in the bibliography) and on the lectures of Grothendieck [11]. I am particularly indebted to W. Neumer and H. G. Tillmann who have respectively read through the first half, and the whole of the manuscript, carefully and critically. M. Landsberg, H. Schaefer and J. Wloka have made important suggestions and observations. Finally I thank the publishers for their speedy and excellent printing. Heidelberg, August 1960 G. Kothe
Preface to the Second Edition The second edition contains a number of corrections, the need for which was kindly pointed out to me by various readers, together with reference to recent articles in which some of the open problems mentioned in the first edition are solved. Apart from this, the text remains unaltered. Frankfurt, October 1965 G. Kothe Preface to the English Edition This English edition is a translation of the second German edition. It differs from the German edition only in several corrections, mainly due to Dr. D. J. H. Garling. I wish to express my sincere gratitude to Dr. Garling for the excellent and careful translation. I am also indebted to Dr. D. Findley for preparing the index. Frankfurt, July 1969 G. Kothe
Contents Chapter One Fundamentals of General Topology § 1. Topological spaces 1 1. The notion of a topological space 1 2. Neighbourhoods 2 3. Bases of neighbourhoods 3 4. Hausdorff spaces 3 5. Some simple topological ideas 4 6. Induced topologies and comparison of topologies. Connectedness . . 4 7. Continuous mappings 6 8. Topological products 7 §2. Nets and filters 9 1. Partially ordered and directed sets 9 2. Zorn's lemma 9 3. Nets in topological spaces 10 4. Filters 11 5. Filters in topological spaces 12 6. Nets and filters in topological products 13 7. Ultrafilters 14 8. Regular spaces 15 §3. Compact spaces and sets 16 1. Definition of compact spaces and sets 16 2. Properties of compact sets 17 3. Tychonoff's theorem 18 4. Other concepts of compactness 18 5. Axioms of countability 19 6. Locally compact spaces 20 7. Normal spaces 22 § 4. Metric spaces 23 1. Definition 23 2. Metric space as a topological space 23 3. Continuity in metric spaces 24 4. Completion of a metric space 25 5. Separable and compact metric spaces 26 6. Baire's theorem 27 7. The topological product of metric spaces 28
X Contents § 5. Uniform spaces 29 1. Definition 29 2. The topology of a uniform space 30 3. Uniform continuity 31 4. Cauchy filters 32 5. The completion of a Hausdorff uniform space 33 6. Compact uniform spaces 35 7. The product of uniform spaces 37 § 6. Real functions on topological spaces 38 1. Upper and lower limits 38 2. Semi-continuous functions 40 3. The least upper bound of a collection of functions 41 4. Continuous functions on normal spaces 42 5. The extension of continuous functions on normal spaces 44 6. Completely regular spaces 44 7. Metrizable uniform spaces 45 8. The complete regularity of uniform spaces 47 Chapter Two Vector Spaces over General Fields § 7. Vector spaces 48 1. Definition of a vector space 48 2. Linear subspaces and quotient spaces 50 3. Bases and complements 50 4. The dimension of a linear space 52 5. Isomorphism, canonical form 53 6. Sums and intersections of subspaces 54 7. Dimension and co-dimension of subspaces 55 8. Products and direct sums of vector spaces 56 9. Lattices , 57 10. The lattice of linear subspaces 58 § 8. Linear mappings and matrices 59 1. Definition and rules of calculation 59 2. The four characteristic spaces of a linear mapping 60 3. Projections 60 4. Inverse mappings 61 5. Representation by matrices 63 6. Rings of matrices 65 7. Change of basis 66 8. Canonical representation of a linear mapping 66 9. Equivalence of mappings and matrices 67 10. The theory of equivalence 68 § 9. The algebraic dual space. Tensor products 69 1. The dual space 69 2. Orthogonality 70 3. The lattice of orthogonally closed subspaces of E* 72
Contents XI 4. The adjoint mapping 73 5. The dimension of E* 74 6. The tensor product of vector spaces 76 7. Linear mappings of tensor products 78 §10. Linearly topologized spaces 82 1. Preliminary remarks 82 2. Linearly topologized spaces 82 3. Dual pairs, weak topologies 85 4. The dual space 86 5. The dual pairs <£*,£> 88 6. Weak convergence and weak completeness 89 7. Quotient spaces and topological complements 90 8. Dual spaces of subspaces and quotient spaces 93 9. Linearly compact spaces 95 10. E* as a linearly compact space 97 11. The topology 2^ 97 12. ^-continuous linear mappings 98 13. Continuous basis and continuous dimension 100 §11. The theory of equations in E and E* 101 1. The duality of E and E* 101 2. The theory of the solutions of column-and row-finite systems of equations 103 3. Formulae for solutions 104 4. The countable case 106 5. An example 107 §12. Locally linearly compact spaces 108 1. The structure of locally linearly compact spaces 108 2. The endomorphisms of \jj 109 3. The theory of equivalence in \jt Ill §13. The linear strong topology 113 1. Linearly bounded subspaces 113 2. The linear strong topology 114 3. The completion 115 4. Topological sums and products 117 5. Spaces of countable degree 119 6. A counterexample 120 7. Further investigations 121 Chapter Three Topological Vector Spaces §14. Normed spaces 123 1. Definition of a normed space 123 2. Norm isomorphism, equivalent norms 125 3. Banach spaces 126 4. Quotient spaces and topological products 127
XII Contents 5. The dual space 128 6. Continuous linear mappings 129 7. The spaces c0, c, ll and /°° 130 8. The spaces /p, 1</?<oo 134 9. (B)-spaces of continuous and holomorphic functions 137 10. The If spaces (p>\) 139 11. The space L00 142 §15. Topological vector spaces 144 1. Definition of a topological vector space 144 2. A second definition 146 3. The completion 148 4. Quotient spaces and topological products 149 5. Finite dimensional topological vector spaces 151 6. Bounded and compact subsets 152 7. Locally compact topological vector spaces 155 8. Topologically complementary spaces 155 9. The dual space, hyperplanes, the spaces LP with 0<p< 1 156 10. Locally bounded spaces, quasi-norms, p-norms 159 11. Metrizable spaces 162 12. The Banach-Schauder theorem and the closed-graph theorem . . 166 13. Equicontinuous mappings, and the theorems of Banach and Banach-Steinhaus 168 14. Bilinear mappings 171 §16. Convex sets 173 1. The convex and absolutely convex cover of a set 173 2. The algebraic boundary of a convex set 176 3. Half-spaces 179 4. Convex bodies and the Minkowski functionals associated with them 180 5. Convex cones 183 6. Hypercones 184 §17. The separation of convex sets. The Hahn-Banach theorem 186 1. The separation theorem 186 2. The Hahn-Banach theorem 188 3. The analytic proof of the Hahn-Banach theorem 189 4. Two consequences of the Hahn-Banach theorem 192 5. Supporting hyperplanes 193 6. The Hahn-Banach theorem for normed spaces. Adjoint mappings . 196 7. The dual space of C(7) 197 Chapter Four Locally Convex Spaces. Fundamentals §18. The definition and simplest properties of locally convex spaces .... 202 1. Definition by neighbourhoods, and by semi-norms 202 2. Metrizable locally convex spaces and (F)-spaces 204 3. Subspaces, quotient spaces and topological products of locally convex spaces 206
Contents XIII 4. The completion of a locally convex space 208 5. The locally convex direct sum of locally convex spaces 211 Locally convex hulls and kernels, inductive and projective limits of locally convex spaces 215 1. The locally convex hull of locally convex spaces 215 2. The inductive limit of vector spaces 217 3. The topological inductive limit of locally convex spaces 220 4. Strict inductive limits 222 5. (LB)-and (LF)-spaces. Completeness 223 6. The locally convex kernel of locally convex spaces 225 7. The projective limit of vector spaces 228 8. The topological projective limit of locally convex spaces 230 9. The representation of a locally convex space as a projective limit . . 231 10. A criterion for completeness 232 Duality 233 1. The existence of continuous linear functional 233 2. Dual pairs and weak topologies 234 3. The duality of closed subspaces 236 4. Duality of mappings 237 5. Duality of complementary spaces 238 6. The convex cover of a compact set 240 7. The separation theorem for compact convex sets 243 8. Polarity 245 9. The polar of a neighbourhood of o 247 10. A representation of locally convex spaces 249 11. Bounded and strongly bounded sets in dual pairs 251 The different topologies on a locally convex space 254 1. The topology 2OT of uniform convergence on 501 254 2. The strong topology 256 3. The original topology of a locally convex space; separability . . . 258 4. The Mackey topology 260 5. The topology of a metrizable space 262 6. The topology 2C of precompact convergence 263 7. Polar topologies 266 8. The topologies Zf and Zlf 267 9. Grothendieck's construction of the completion 269 10. The Banach-Diedonne theorem 272 11. Real and complex locally convex spaces 273 The determination of various dual spaces and their topologies .... 275 1. The dual of subspaces and quotient spaces 275 2. The topologies of subspaces, quotient spaces and their duals . . . . 276 3. Subspaces and quotient spaces of normed spaces 279 4. The quotient spaces of Z1 280 5. The duality of topological products and locally convex direct sums . 283 6. The duality of locally convex hulls and kernels 288 7. Topologies on locally convex hulls and kernels 291
XIV Contents Chapter Five Topological and Geometrical Properties of Locally Convex Spaces §23. The bidual space. Semi-reflexivity and reflexivity 295 1. Quasi-completeness 295 2. The bidual space 297 3. Semi-reflexivity 298 4. The topologies on the bidual 300 5. Reflexivity 302 6. The relationship between semi-reflexivity and reflexivity 304 7. Distinguished spaces 306 8. The dual of a semi-reflexive space 307 9. Polar reflexivity 308 §24. Some results on compact and on convex sets 310 1. The theorems of Smulian and Kaplansky 310 2. Eberlein's theorem 313 3. Further criteria for weak compactness 315 4. Convex sets in spaces which are not semi-reflexive. The theorems ofKLEE 319 5. Krein's theorem 323 6. Ptak's theorem 326 §25. Extreme points and extreme rays of convex sets 330 1. The Krein-Milman theorem 330 2. Examples and applications 333 3. Variants of the Krein-Milman theorem 336 4. The extreme rays of a cone 337 5. Locally compact convex sets 339 § 26. Metric properties of normed spaces 342 1. Strict convexity 342 2. Shortest distance 343 3. Points of smoothness 345 4. Weak differentiability of the norm 347 5. Examples 350 6. Uniform convexity 353 7. The uniform convexity of the lp and LP spaces 355 8. Further examples 359 9. Invariance under topological isomorphisms 360 10. Uniform smoothness and strong differentiability of the norm . . . 363 11. Further ideas 366 Chapter Six Some Special Classes of Locally Convex Spaces § 27. Barrelled spaces and Montel spaces 367 1. Quasi-barrelled spaces and barrelled spaces 367 2. (M)-spaces and (FM)-spaces 369
Contents XV 3. The space//(©) 372 4. (M)-spaces of locally holomorphic functions 375 § 28. Bornological spaces 379 1. Definition 379 2. The structure of bornological spaces 380 3. Local convergence. Sequentially continuous mappings 382 4. Hereditary properties 383 5. The dual, and the topology <XCo 384 6. Boundedly closed spaces 386 7. Reflexivity and completeness 388 8. The Mackey-Ulam theorem 389 §29. (F)- and (DF)-spaces 392 1. Fundamental sequences of bounded sets. Metrizability 392 2. Thebidual 394 3. (DF)-spaces 396 4. Bornological (DF)-spaces 399 5. Hereditary properties of (DF)-spaces 401 6. Further results, and open questions 403 § 30. Perfect spaces 405 1. The a-dual. Examples 405 2. The normal topology of a sequence space 407 3. Sums and products of sequence spaces 409 4. Unions and intersections of sequence spaces 410 5. Topological properties of sequence spaces 412 6. Compact subsets of a perfect space 415 7. Barrelled spaces and (M)-spaces 417 8. Echelon and co-echelon spaces 419 9. Co-echelon spaces of type (M) 421 10. Further investigations into sequence spaces 423 §31. Counterexamples 424 1. The dual of/00 424 2. Subspaces of /°° and Z1 with no topological complements 426 3. The problem of complements in lp and LP 428 4. Complements in (F)-spaces 431 5. An (FM)-space 433 6. An (LB)-space which is not complete 434 7. An (F)-space which is not distinguished 435 Bibliography 437 Author and Subject Index 447
CHAPTER ONE Fundamentals of General Topology In this preliminary chapter we gather together those ideas and theorems of general topology which we shall need later. We have also given proofs of the theorems, since an understanding of the methods of topology is essential for the study of vector spaces. For detailied information one must of course refer to texts on general topology; we mention Bourbaki [5], Kelley [2], Lefschetz [1] and Schubert [1]. The account given here follows Bourbaki closely. § 1. Topological spaces 1. The notion of a topological space. A topology % is defined on a set R when a class O of subsets of R is given, which satisfies the conditions: (Ol) K and the empty set are in O; (O 2) O contains with every finite collection of sets their intersection, and with every arbitrary collection of sets their union. The sets of O are called the open sets of R. A set R with a topology X defined on it is called a topological space. The elements of R are called the points of the space. A subclass 33 of O is called a basis of open sets of R if every open set is a union of sets of 33. A subclass of O is called a sub-basis when the finite intersections of its sets form a basis. The topology X is determined by a basis or sub-basis of open sets. The complement R~0 of an open set 0 is called a closed set of R. The class 21 of all closed sets of R clearly has the properties: (A 1) R and the empty set are in 21; (A 2) 21 contains with every finite collection of sets their union, and with every arbitrary collection of sets their intersection. As a result, a topology on a set R can also be defined by giving a class 21 with the properties (Al) and (A2). In this case the open sets are the complements of the closed sets. Bases and sub-bases of 21 are defined as above, exchanging the notions of "union" and "intersection". When we speak of a basis of R, however, we shall always mean a basis of open sets of R. 1 Kothe, Topological Vector Spaces 1
2 § 1. Topological spaces 2. Neighbourhoods. A third way of introducing a topology is by giving the collection of all neighbourhoods. A subset of the topological space R which contains an open set containing the point x is called a neighbourhood of x. Let Sft(x) be the class of all neighbourhoods of x. It is easy to confirm the following properties of91(x): (N 1) $1 (x) is non-empty and x belongs to each set of^l(x); (N2) If a set belongs to %l(x) then so does every larger subset of R; (N 3) The intersection of a finite collection of sets ofSH{x) lies in 9l(x); (N4) For every U in9l(x) there is a V in$l(x) such that Ue9l(y) for each y in V. For (N4) we observe that every open neighbourhood V of x contained in U has the required property. Conversely, suppose that for each x in a set R a non-empty class 9l(x) of subsets of R is given, and that (N 1) to (N4) are satisfied. If 9l(x) is to be, for each x, the class of all neighbourhoods of x in some topology X on R, then the non-empty open sets must be identical with those subsets 0 of R for which Oe^l(x) whenever xeO. The class O of all these sets O together with the empty set satisfies (01) and (02). For the empty set is in O, and by (N2) so is R; by (N3) the intersection of finitely many sets 0 is again a set of O, and by (N2) so is an arbitrary union of sets 0. Thus O defines a topology %. We still have to show that the ^-neighbourhoods of x coincide with the sets of 9l(x). Every ^-neighbourhood of x contains a set 0 which belongs to 91(x), so that by (N2) every ^-neighbourhood of x is in 9l(x). Conversely, suppose that U belongs to 9l(x). We consider the subset U1 of all y in U for which Ue9l(y). Since x is in U1 it is enough to show that Ux belongs to O. By (N4) there is for each y in U a V in 9l(y) for which Ueyi(z) for each z in V. From the definition of Uu z lies in Uu so that Kcz U1 and so, by (U2), Ux e9i(y). Thus we have shown (1) Suppose that for each x in a set R a class ^l(x) of subsets is given, and that (N1) to (N4) are satisfied. Then there is a unique topology on R for which 91 (x) is the class of all neighbourhoods of x, for each x in R. Two topologies % and %' on a set R thus give rise to the same topological space when they give either the same open sets or the same closed sets or the same neighbourhoods of each point. Two topological spaces Rx and R2 are homeomorphic when there is a one-one mapping of the points of Rr onto the points of R2 which sends every open set of Rx into an open set of R2, and conversely. Such a mapping is called a homeomorphism. Closed sets or the classes 9l(x) can be used instead of open sets in the definition of homeomorphism.
4. Hausdorff spaces 3 3. Bases of neighbourhoods. If91(x) is the class of all neighbourhoods of a point x of the topological space R, a subclass 35(x) of 9i(x) is called a base of neighbourhoods of x (or fundamental system of neighbourhoods of x) if every neighbourhood in $l(x) contains one in 35(x); to put it another way, $l(x) is obtained from 35(x) by taking all those subsets of R which contain some set in 35(x). If a base of neighbourhoods 3S(x) is given for each x, we speak of a base 35 of neighbourhoods in R. For a base 35 of neighbourhoods consisting solely of open sets Hausdorff's three axioms follows easily from (Nl) to (N4), together with the characterization of open sets given in 2.: (H 1) Every point x has at least one neighbourhood in 35 (x), and lies in each of its neighbourhoods; (H2) The intersection of two neighbourhoods in 3S(x) contains a neighbourhood in 35 (x); (H3) If y lies in Ke35(x), there is a WeW(y) with W^V. Here also the converse holds, that a unique topology is defined by a base 35 of neighbourhoods in R which satisfies (H 1) to (H3). This follows, since we obtain a class of neighbourhoods satisfying (Nl) to (N4) by taking as neighbourhoods all subsets of R larger than the given neighbourhoods. Thus we have a fourth method of introducing a topology on a set R. Starting from a base of neighbourhoods, a set is open if and only if whenever it contains a point it contains a basic neighbourhood of the point. The open sets of a topological space other than the empty set always form a base of neighbourhoods. Two base of neighbourhoods 35 and 35' on the same set R are called equivalent when they define the same topology. This is obviously the case if and only if the classes of all neighbourhoods determined by them are the same. From this, Hausdorff's criterion, which will often be used later, follows immediately: (1) Two bases of neighbourhoods 35 and 35' on the same set R are equivalent if and only if for each x in R every neighbourhood of x in either base always contains a neighbourhood of x in the other base. 4. Hausdorff spaces. A topological space R is said to be a Hausdorff space, or separated, if it satisfies the fourth of Hausdorff's axioms: (H4) Any two distinct points of R possess neighbourhoods in 35 without common points. This can also be expressed in the following way (T2) Two distinct points of R always lie in disjoint open sets. T2 is often called the second, or Hausdorff, separation axiom, and Hausdorff spaces are called T2-spaces (cf. Lefschetz [13]). i*
4 § 1. Topological spaces If R is a general set, and all the subsets of R containing x are taken as neighbourhoods of x, for each x in R, then R, with the discrete topology defined in this way, is a Hausdorff space. 5. Some simple topological ideas. A point x is called an interior point of a subset M of a topological space, if a whole neighbourhood of x lies in M. The collection of all interior points of M forms an open set, the interior of M. A point x is called an exterior point of M if it is an interior point of the complement R~M. A set U => N, where Af is an open set containing M, is called a neighbourhood ofM. A point x is called a closure point of the set M if every neighbourhood of x contains at least one point of M. The set of all closure points of a set M is called the closure MofM. Since the complement of M is an open set, M is a closed set, and indeed is the intersection of all the closed subsets of R which contain M. Thu|a set is closed if and only if it coincides with its closure. In particular M = M, for any set M. A point x is called an accumulation point of the set M if every neighbourhood of x contains at least one point of M distinct from x. A closure point of M fails to be an accumulation point of M if and only if it is an isolated point of M—i.e. a point which has a neighbourhood containing no other point of M. Clearly M is closed if and only if it contains all its accumulation points. The boundary of a set M is the intersection of the closures of M and R~M. A boundary point ofM is thus a closure point of M and R~M. Every closed set contains its boundary and every open set is disjoint from its boundary. The set N is said to be dense in M when M cz N, everywhere dense when N = R, and nowhere dense when N has no interior points. The boundary of any open or closed set is nowhere dense. As an application of these ideas we show (1) In a Hausdorff space the intersection of the closed neighbourhoods of a point contains the point alone. If x0 is the given point and y is a point different from x0 then by (H4) there exist a neighbourhood U(x0) and a neighbourhood V(y) with U nV empty. But then y is an interior point of R~U, and so an exterior point of (/, and thus y does not lie in the closure of U. From (1) there follows immediately (2) The only Hausdorff topology on a finite set is the discrete topology. 6. Induced topologies and comparison of topologies. Connectedness. If S is a subset of the topological space R, the topology % of R induces a topology on S when the sets SnO, 0 open in R, are taken as open sets in S. The induced topology is also obtained by considering the inter-
6. Induced topologies and comparison of topologies. Connectedness 5 sections of the closed sets with S or the intersections with S of neighbourhoods of the points of S. Notice that a set which is open or closed in S need not be so in R. If R is Hausdorff, then so is S in the induced topology, which in general we shall denote by X again. If two topologies X{ and X2 are defined on a set R, X1 is said to be finer (or stronger) than X2 if every ^-neighbourhood is also a Xi-neighbourhood; X2 is said to be coarser (or weaker) than 3^. That Xx is finer than X2 can also be expressed by saying that the class Oj of 3^-open sets includes the class C)2 °f ^2 open sets. The same holds for the classes of closed sets; for this reason we write Xx => X2. If Xl is finer than X2, every 3^-closure point of a set M is also a 32-closure point, but not conversely; in general a smaller set is obtained by forming the Xi-closure than by forming the 32-closure. If finitely or infinitely many topologies Xa are defined on a set R, there is a finest topology X among the topologies on R which are coarser than every Xa: the ^-neighbourhoods of a point x are those sets which are ^-neighbourhoods of x for each a. If Da are the classes of ^-open- sets, then the class O of 2-open sets satisfies O = (°)©a, since O ^ Oa a for each a. X is called the intersection of the Xa. X need not be Hausdorff, even if the Xa are Hausdorff. If f] Oa consists only of the empty set and R, then X is the trivial topology, in which every point has the single neighbourhood R. Similarly, given Xa9 there is a coarsest topology X among the topologies which are finer than every Xa. This is called the union of the Xa. A base of ^-neighbourhoods of x in R is formed by the ^-neighbourhoods of x, for all a, together with their finite intersections. The discrete topology (4.) can occur here as an extreme case. The union of Hausdorff Xa is again Hausdorff. A topological space R is called connected when it is not the union of two non-empty disjoint open sets. This is equivalent to saying that R is not the union of two non-empty disjoint closed sets, or that R contains no proper non-empty sets which are both open and closed. A subset S of R is said to be connected, when S is connected as a topological space with the induced topology. R is connected whenever every two points of R lie in a connected subset. If R could be divided into two non-empty open sets Rx and R2, than any subset S containing two points from Rx and R2 would divide into two non-empty open subsets RtnS and R2nS of S. We denote by P the field of real numbers, and by P" real n-dimensional space in the natural (Euclidean) topology.
6 § 1. Topological spaces Since the straight line joining two points of P" is connected, n-di- mensional space P" is also connected. A topological space is called totally disconnected when it has no connected subsets other than the one-point sets. Every discrete space is totally disconnected, but not conversely; for example the rational numbers form a totally disconnected space in the topology induced by the natural topology of P. 7. Continuous mappings. Let A be a mapping from the topological space #! into the topological space R2—that is, an assignment which sends each xeR{ to AxeR2. We also speak of a function on Ri with values in R2, although this term will generally only be used when R2 is a space of numbers. Every such point-mapping gives rise to a mapping from the class of subsets of Rt into the class of subsets of R2, which will again be denoted by A. In detail, if M is a subset of Rl9 the set of all Ax, xeM, forms a subset A(M) of R2, which is called the image set or image of M. In particular AiR^ is called the image space of A. If A is one-one and y = Ax, the correspondence A{~l)y = x defines a one-one mapping A{~1) from AiR^ onto Rx. We call A{~1) the inverse of A. This also gives rise to a mapping A{~1) from the class of subsets of A(R{) onto the class of subsets of ^^ If A is not one-one, the point-mapping has no inverse. To every subset N of ;4(#i), however, we can make correspond its inverse image Ai~1)(N), the set of all x in Kt with AxeN. In this way, for every mapping A the inverse A{~1) is defined as a mapping of the class of subsets of AiR^ into the class of subsets of Ru If M is a general subset of R2, by A{~l)(M) we shall always mean Ai-ViMnAiRJ). If a mapping A from Rt into R2 maps every open (respectively closed) subset of #! into a set which is open (respectively closed) in ^(ftj (but not necessarily open or closed in R2\), then A is called an open (respectively closed) mapping. In the same way the inverse A{~1) is said to be open (closed) when every open (closed) subset of AiR^ has an open (closed) inverse image. A mapping A from #! into R2 is said to be continuous at x0, when for each neighbourhood V of Ax0 in R2 there exists a neighbourhood U of x0 in #! whose image lies entirely in V. Clearly we can restrict attention to neighbourhoods lying in a given base of neighbourhoods. If A is continuous at every point x of Rl9 A is said to be continuous (on RJ. (1) The following properties of A are equivalent: a) A is continuous, b) A{~1) is open, c) A{~ l) is closed.
8. Topological products 7 Proof. If A{~1} is open, then for a given open neighbourhood V of Ax0 the set A{~l)(V) is open, and so is a neighbourhood of x0, whose image under the mapping A is contained in V. If conversely A is continuous and M is a subset of A(RX) which is open in AiR^ then if Ax0eM there is a whole neighbourhood contained in M. The image of this neighbourhood under the mapping A{~1) contains a neighbourhood of x0, so that A{~l)(M) is open in Rl. Since A{~1) maps A(Rt)~M into Rl^Ai~l)(M), A{~1) is open if and only if it is closed. If the continuous mapping A is one-one, A{~1) need not be continuous. (2) A one-one mapping A from Rl onto R2 is a homeomorphism if and only if A and A{~1) are continuous, and if and only if A and A{~1) are open (closed). Under a homeomorphism there is a one-one correspondence between the neighbourhoods, the open sets and the closed sets of Rt and those of R2. On the other hand, it follows by (1) from the continuity of A and A{~1] that A{~1] and A are open—i.e. the open sets of R{ and R2 are in one-one correspondence, and A is a homeomorphism. (3) A mapping A is continuous if and only if it sends every closure point of a set into a closure point of the image set. Proof. It follows directly from the definition of continuity that a closure point is sent into a closure point of the image set. We remark that the image of an accumulation point need not be an accumulation point. The other part of (3) does however hold for accumulation points: if A is not continuous, there exists at least one neighbourhood V of a point Ax0 for which points xv can be found in every neighbourhood U of x0, whose images do not belong to V. The set of these xv has x0 as accumulation point, and A x0 is not a accumulation point of the set of images Axv. The composition of finitely many continuous mappings is always a continuous mapping. 8. Topological products. If Ri9...,Rn are given sets, the set R of n all w-tuples x = (x1}...,xn), xteRh is denoted by Rl x ••• x Rn = TT Rf. i = 1 If the Rt are topological spaces, R becomes the topological product n of the R( when the class of all sets U = TT L/f(xf) is taken as base of i= 1 neighbourhoods of x, where L^x,) is a general neighbourhood of x, n in R(. We use the expression TT R( for these topological spaces as well. i=l
8 § 1. Topological spaces These definitions can be extended to arbitrarily many factors. If Ra, aeA, are given, R = TT Ra denotes the set of all functions x(a) = xaeR0i. aeA R is called the set-theoretic product or Cartesian product of the Ra. If the Ra are topological spaces, R becomes the topological product of the R^ when all subsets U = TT W, are taken as base of aeA a neighbpurhoods of the point x, where W^R* for all but finitely many a, and Wp=Up(xp) for the others, where Up(xp) is a general neighbourhood of xp in Rp. If Ra = S for all a, we write SA for the topological product; in particular we write Sn when there are n equal factors and S03 when there are countably many. Thus if P is the set of real numbers with its natural topology, P" is n-dimensional space. Pw is the space of all sequences, with the topology which has just been defined. By the parallelotope ^3A we mean the topological product SA, where S is the closed interval [0,1]. If the Ra are Hausdorff, TT Ra is also Hausdorff. aeA The mapping Pa, which sends each xeR to its component xaeKa, is called the projection of R onto Ra. It is a continuous mapping from R onto Ka, and the topology of R is the coarsest topology for which all the projections Pa are continuous. For if Pa is continuous, the inverse image of Ua{xa) must be a neighbourhood of x in R, by 7.(1). By taking finite intersections of these inverse images we obtain all the neighbourhoods in the given base of neighbourhoods. The projection Pa is open, since an open set contains a neighbourhood TT W% of each point, the projection contains the neighbourhood W% of the image, and so the a-components of the points of an open set form an open set. Pa need not be closed, as is shown for example by considering the closed set of all (n,\/n), h = 1,2,..., in P2. (1) If Ma are subsets of Ra, the closure of the set TT Ma is TT Ma, where Ma is the closure of Ma in Ka. Proof. It is immediately clear that at least one element of TTMa lies in each of the neighbourhoods belonging to the given base of neighbourhoods of xeTTMa, and conversely only elements of TTMa can be closure points. // the Ma are all closed, so is TTMa. If the Ma are open, TTMa need not be open, when A has infinitely many elements. If A is a mapping from the topological product RtxR2 into the topological space S which is continuous at the point {x[0\x{2}), the mapping x1-^A(xl,x{2)), from Rt into S is continuous at the point x(10) (partial continuity).
2. Zorn's Lemma 9 § 2. Nets and filters 1. Partially ordered and directed sets. A set H is said to be partially ordered or semi-ordered if a relation x^y (x less than or equal to y) is defined for certain pairs of its elements, which is reflexive (x^x), transitive (x^y and y^z imply x^z) and antisymmetric (x^y and y^x imply x = y). For x^y we also write y^x; x<y means xf^y and x + y. A partially ordered set is called totally ordered or simply ordered if one of the relations x^y or y^x always holds for any two of its elements x and y. A partially ordered set H is called a directed set when for any two elements x and y there always exists zeH for which x^z and yrgz. H is said to be inversely directed when for any two elements x and y there always exists zeH for which z^x and z^y. Every totally ordered set is both directed and inversely directed. If x is a point of a topological space, the neighbourhoods of a base of neighbourhoods of x form a directed set under the set-theoretic relation =>. Let M be a subset of the partially ordered set H. It is partially ordered under ^. M is said to be bounded above (bounded below) if there exists yeH for which x^y (y^x) for all xeM. y is called an upper (lower) bound ofM. Every finite subset of a directed set is bounded above. If the set of upper (lower) bounds of M has a least (greatest) element y0, y0 is called the least upper bound (greatest lower bound) of M. M is called a maximal (minimal) element of M if there exists no x in M with z<x (x<z). A least element of M is always minimal, but not conversely. 2. Zorn's Lemma. A totally ordered set is said to be well- ordered if each of its non-empty sets has a least element. We shall take the results of classical set theory for granted (cf. Hausdorff [2] and Kamke [1], for example). In particular, we shall assume the validity of the axiom of choice. Every set can then be well- ordered, using the ordinals as index-set. We shall also make occasional use of transfinite induction and the theory of cardinal numbers. As an example of transfinite induction, we mention Zorn's lemma (1) If every totally ordered subset of a partially ordered set H has an upper bound, H has at least one maximal element. Proof. Let xa, a = 0,1,..., be a well-ordering of the elements of H. We determine a totally ordered subset G of H by transfinite induction: x0 belongs to G; if it has been determined which xp belongs to G, for
10 § 2. Nets and filters all P<y, then xy belongs to G if and only if xp<xy for all xpeG. By hypothesis G has an upper bound z. Since z is an xa, z must lie in G and so must be the largest element of G and a maximal element of H. Zorn's lemma is so general that most applications of the well- ordering principle are special cases of this result, so that we shall not need to make repeated use of the well-ordering principle. As a special case, for subsets of a set M partially ordered by <= we have (2) // H is a class of subsets of a set M with the property that if a collection of subsets in H is totally ordered by cz, then its union also belongs to H, then H has at least one maximal subset. Let us remark that Zorn's lemma can be deduced diretly from the axiom of choice, without the aid of the well-ordering principle (cf. Kamke [1], for example). In fact the axiom of choice, the well-ordering principles and Zorn's lemma are equivalent assumptions (cf. Birkhoff [3] or Hermes [1] as well). 3. Nets in topological spaces. If x„ is a sequence of points in a topological space R, xn is said to be convergent to x0eR if for each neighbourhood U of x0 there exists n0(U) such that xneU for all n^n0(U). x0 is called limit of the sequence xn, and we write x„->x0. If R is Hausdorff, a sequence can only have one limit. In this case we write x0 = \imxn. n In a general topological space R an accumulation point of a set M need not be limit of a sequence of points of M. The parallelotope S$A, with uncountable A, gives an example of this (cf. § 3, 4.). By using directed sets, however, the concept of limit is generalized in such a way that every accumulation point becomes a limit. Let A be a directed set, and let M be a general set. If for each aeA an xa e M is given, the xa form a net inM. When A = 1,2,... we obtain sequences as.special cases of nets. The net xa, aeA, is said to be convergent to x0eR if for each neighbourhood U of x0 there exists (S(U)eA such that xyeU for all y^P(U). xo ls called limit of xa, and we write xa->x0. Again, the limit of a net in a Hausdorff space must be unique, and this property characterizes Hausdorff spaces (proof!). In this case we also write x0 = lim xa. a The concept of subsequence can also be generalized: a subset B of a directed set A is cofinal if for each aeA there exists /7eB with /?g^a. If a subset of a directed set is not cofinal, then its complement is. If xa, aeA, is a net, the xp form a cofinal subnet if the f} form a cofinal subset of A. If x0 is limit of the net xa, x0 is also limit of every cofinal subnet. A point y0 is called an adherent point of the net xa if every neigh-
4. Filters 11 bourhood of y0 contains a cofinal subnet. Every adherent point of the net xa is a closure point of the subset of R consisting of the distinct xa, but not conversely. Every limit of xa is an adherent point. From this follows one half of (1) A subset M of a topological space R is closed if and only if it contains the limits of all the convergent nets of elements of M. On the other hand if a is a closure point of M, an xveM can be picked out of every neighbourhood U of a base of neighbourhoods of a. The sets U form a directed system under id, by 1. The net xv clearly has limit a. (2) A mapping A is continuous at x0 if and only if xa->x0 always implies Axa-^Ax0. Necessity follows immediately from the definition of continuity of A and the definition of convergence of a net. Conversely, suppose that A is not continuous at x0, and that U={U} is a base of neighbourhoods of x0. Then there exists a neighbourhood V of Ax0 and an xv in each UeVL such that Axv does not lie in V. But then the xv form a net with xv^x0, and the image net Axv does not converge to Ax0. We remark that we obtain a proof of continuity at x0 using nets xv, where U runs through a fixed base of neighbourhoods of x0. An analogous remark applies to (1). If Ax„-+Ax0 for all sequences x„->x0, we call A sequentially continuous at x0. Sequential continuity of A does not in general imply continuity of A. 4. Filters. Closely related to the concept of net is the concept of filter. A non-empty class g = {Fa} of subsets of a set M is called a filter on M if (F1) Every subset of M containing an Fa belongs to g ; (F2) The intersection of finitely many Fa belongs to g; (F3) The empty set does not belong to g. The class of all subsets of the natural numbers with finite complements forms a filter. More generally a filter on M is obtained from a net xa, aeA, xaeM, by forming the set Fa of all distinct xp with /?^a, for each a, and collecting these sets and the subsets of M containing them into a class g. We call this filter the filter corresponding to the net xa. Conversely the sets Fa of a filter g form a directed set under =>; moreover we obtain a net xa if we choose an xa from each Fa, and order the a by
12 § 2. Nets and filters the partial order of the Fa which has just been described. Nets formed in this way are called the nets corresponding to a filter. An especially important example is given by the filter which consists of all the neighbourhoods of a point x0 of a topological space, the neighbourhood filter of x0. A non-empty subclass 33 of a filter g on M is called a filter-base of g if it satisfies the conditions (B 1) The intersection of two sets of 33 contains a set of 33; (B 2) The empty set does not belong to 33, and if g consists of all those subsets of M which contain a set of 33. Conversely, given a collection 33 of subsets of a set M which satisfies (Bl) and (B2), a filter is obtained by taking all the larger sets. Because of this, such a collection is called a filter-base. In this context a base of neighbourhoods of a point is nothing else than a filter-base of the filter of all neighbourhoods of the point. Hausdorffs criterion in § 1, 3. generalizes to (1) Two filter-bases 33 and 33' define the same filter if and only if every set of either base contains a set of the other base. Such bases are called equivalent. A collection 6 of subsets of a set M, for which every finite collection of sets has a non-empty intersection, gives rise to a filter-base, by taking all these finite intersections. Such a collection is called a sub-base of the filter which it determines. If g and g' are two filters on the same set M, and if g <= g', that is if g is a subclass of g', then g is said to be coarser than g', and g' finer than g. If %' is a finer topology on M than % then the neighbourhood filter of a point x0 relative to 3/ is finer than that relative to X. If MczAT and g = {Fa} is a filter on M, then the Fa form the base of a filter on N which we shall in general denote by g again. Conversely if g = {Fa} is a filter on N, then, provided that all of them are non-empty, the sets FanM form a filter on M, the restriction of g to M. 5. Filters in topological spaces. Following the pattern of 3. we make the following definition. A filter g = {Fa} on a topological space R converges to x0 if there exists an Fpa U for every neighbourhood U of Xq. Xq is called limit of the filter, and we write Fa-+x0. If R is Haus- dorff, there is at most one limit, and then we write limg = x0 or limFa = x0. If g converges to x0, so does every finer filter g'. a If the filter is given by a basis {£a}, the condition for convergence reads: every U(x0) must contain a Ba. A point x0 which is a closure point of all the Fa is called an adherent point of the filter. For this it is sufficient for x0 to be a closure
6. Nets and filters in topological products 13 point of all the sets of a base of the filter. If g has limit x0 in a Haus- dorff space, x0 is the unique adherent point of g. The converse is not true, as the filter on P with base Fn = {0} u [n, oo), n = 1,2,..., shows. (1) // x0 is an adherent point of g={Fa} and {Up} is the neighbourhood filter of x0, then all the Fan Up form a filter convergent to x0, which is finer than g. The following relations between these ideas and the corresponding ones for nets are easily established: (2) The filter {Fa} has x0 as limit if and only if xa->x0 for every net corresponding to it. The net xa has limit x0 if and only if the corresponding filter has limit x0. (3) // xa is a net corresponding to {Fa}, the filter corresponding to the net xa is finer than {Fa}, and has exactly the same adherent points as xa. The adherent points of the nets corresponding to {Fa} are thus adherent points of {Fa}. We remark that conversely an adherent point of {Fa} need not always be an adherent point of every corresponding net. Theorems (2) and (3) also hold for filter-bases {Bp} and the corresponding nets xpeBp. If M is a subset of R and x0 is a closure point of M, consideration of the filter {Mnl^}, where {Up} is the neighbourhood filter of x0 in R, shows the following, the direct analogue of 3.(1): (4) A subset M of a topological space R is closed if and only if it contains all the limits in R of filters on M. If A is a mapping of the set Ml into the set M2 and g={Fa} is a filter on M1? then since A(FJ is non-empty and A(FanFp) cz A(FjnA(Fp) the image sets A(FJ form the base of a filter on A(MX), and also on M2, which we call the image-filter i4(g). The images of a base also form a base of the image-filter. If (& = {Gp} is a filter on A{Ml\ the sets ^(_1)(^/?) generate a filter on M1? which we call the inverse-image filter A{-1)(($>). From 3.(2) and (2), or directly, we have (5) A mapping A from a topological space Rx into a topological space R2 is continuous at x0 if and only if Fa->x0 always implies that AFa-+Ax0. 6. Nets and filters in topological products. Let R= TT Rp be a topological product. PeB (1) A net x{cc)eR has x{0) as limit if and only if x^a)->x^0), for each p. Necessity follows from the continuity of the projections Pp of R onto Rp (§ 1, 8.) and 3.(2); sufficiency results from the fact that a neigh-
14 § 2. Nets and filters bourhood T7 WB of x{0) has only finitely many WB + RB, so that an in- dex y exists for which xfeWp{x^\ for all p and all <S^y; i.e. x(<5)eT7H^. If ^4 is a mapping of the topological space S into the product R, then PpAx = Apx is a mapping of S into Rp. From (1), we have (2) The mapping A of a topological space S into R = T\ Rp is continuous at x0 if and only if all the Ap are continuous at x0. Let g={Fa} be a filter on R = T\ Rp.The projections of the elements of Fa on Rp form a set F*p. The filter Pp(%) = %p with sets F*p is called the projection of the filter g on /^. (3) T/ze filter g converges to x{0) on R if and only if every projection g^ converges to xft\ By 5.(2), this is just another wording of (1). If a filter g^ is given on each Rp, the p r o d u c t - f i 11 e r TT g^ is defined as the filter g on T\Rp which has as base all sets T\A09 where Ap = Rp for all but finitely many p, and Ap is an arbitrary set of g^ for finitely many p. The product 1733^ of filter-bases of Fp is likewise a base for TTg^. In this setting, the neighbourhood filter of a point x ofTTi^ is the product of the neighbourhood filters of its components. 7. Ultrafilters. The filters on M form a partially ordered set under cz. If a collection ga of filters is given, P)ga is again a filter, since (Ft), (F2) a and (F3) are satisfied and M always belongs to it. P)ga is the greatest a lower bound of the filters ga. The union (Jga forms a filter (the a least upper bound of the ga) if and only if the intersection of finitely many sets from distinct ga is never empty. If {ga} is a totally ordered collection of filters on M, the least upper bound (Jga exists. Using Zorn's lemma [2.(1)] we consequently have a (1) For every filter g on M there exists a finer maximal filter, a so- called ultrafilter. (2) A filter g is an ultrafilter on M if and only if the following holds for any two subsets A and B of M: if A u£eg, then g contains at least one of the two sets A and B. Proof. If g does not contain A and B, all the subsets N of M with NkjAe^ form a filter, since (IV1ni\[2)u/l = (]V1u/l)n(]V2u/l) + /l, so that Nt n N2 is non-empty. This filter is then finer than g, since the set B has been added to the sets of g. On the other hand, if the condition is satisfied g contains at least one of every pair of sets A and M ~A.
8. Regular spaces 15 If there were a finer filter, then it would have to contain some A and its complement M~A, and so would also contain the empty set An(M~A). (3) The image of an ultrafilter is again an ultrafilter. Let g = {F^} be an ultrafilter on M, where A maps M into N. If A(FP) were not the base of an ultrafilter on JV, then by 4.(1) there would be a finer filter © = {Gy} on N, with at least one Gyo c 4(M) containing no A(FP). The filter defined by A{~l)(Gyo) and g on M would then be finer than g, since y4(_1)(GVo) could contain no F/j. If g is an ultrafilter on M czN, then by (3) 5 defines an ultrafilter on N. Equally, by (2) the restriction to M of an ultrafilter on N is again an ultrafilter; the restriction exists if and only if M belongs to the ultra- filter. (4) In a topological space an ultrafilter g is either convergent or has no adherent points. If g nas an adherent point x0, then by 5.(1) there is a finer filter convergent to x0. Since g is maximal, this filter coincides with g. (5) 4 mapping A from a topological space Rx into a topological space R2 is continuous at x0 if and only if for every ultrafilter g = {Fp}, Fp^x0 always implies that A(Fp)->A{x0). Proof. By 5.(5) it is enough to show that, given a filter © = {G^} convergent to x0 but with an image-filter which is not convergent to Ax0, there exists an ultrafilter with the same property. Since A(GI}) does not converge to Ax0, there is a neighbourhood V of A x0 for which all the sets Hp = (R2 ~ V) n A{GP) are non-empty. But then the sets Mp = A{-l\H[i)nG(i form the base of a filter 9W which is finer than ©. Any ultrafilter g which is even finer is convergent to x0, and its image 4(g) is an ultrafilter which does not converge to A(x0), since the sets A(MP) = HP of the filter are disjoint from V. 8. Regular spaces. A Hausdorff space R is said to be regular if it satisfies the condition (R) The closed neighbourhoods of each point form a base for the neighbourhood filter. If 95(x) is a base of neighbourhoods of x, and U is an arbitrary closed neighbourhood of x, then there exists Ve%$(x) with Vcz U, and so also Kc= U: The closures of the neighbourhoods of a base of neighbourhoods of a regular space form a base of neighbourhoods for R. Condition (R) is equivalent to the condition (R') // M is closed in R and x does not belong to M, then disjoint neighbourhoods of M and x can always be found.
16 § 3. Compact spaces and sets The simple proof is left to the reader. In regular spaces Fa-»x0 always implies that Fa-»x0. (1) Every subspace S of a regular space R is regular. For S is Hausdorff and the intersections with S of the closed neighbourhoods in R of a point x of S are closed in S, and form a base of neighbourhoods in S. (2), The topological product of regular spaces is regular. In certain circumstances continuous mappings into regular spaces can be extended. Let M be dense in the topological space Rx. Let A be a continuous mapping from M into the regular space R2. If x0 is a point of R{ which is not in M and (/'a=[/anM, where {Ua} is the neighbourhood filter of x0 in Rl9 then A can be extended continuously to x0 only if A{U'J converges to a point of R2, which we define to be the image Ax0. This requirement is stronger than that of continuity of A on M. It can also be expressed in terms of nets: if x0eRl9 then for all nets xa->x0, with xaeM, the image nets Ax^ must all converge to one and the same element of R2. We still have to show that the mapping which has now been defined on the whole of Rx is continuous at each point. Let V be a neighbourhood of Ax0. Because R2 is assumed to be regular we can suppose that V is closed. There exists an open neighbourhood U of x0 with A(U n M) c V. If y is an arbitrary point of U, there is a net ya in UnM with ya-+y, so that Ay is limit of the net Aya9 which lies in A(U nM)czV. Thus Ay is in P= K, and A(U)czV. Thus we have established (3) A continuous mapping A from a dense subset M of a topological space Rx into a regular space R2 can be continuously extended to the whole of Rx provided that for each point xgRx with neighbourhood filter {£/a} A(Uar\M) always converges. The extension is clearly unique. § 3. Compact spaces and sets 1. Definition of compact spaces and sets. A Hausdorff space R is said to be compact if every filter on R has at least one adherent point. Since every filter is contained in an ultrafilter, by §2,7.(1), we can also say (1) R is compact if and only if every ultrafilter on R is convergent. (2) A Hausdorff space R is compact if and only if every collection {Aa} of closed subsets Aa with empty intersection f^A^ contains a finite collection of subsets with an empty intersection. a
2. Properties of compact sets 17 Proof. Suppose that R is compact, that f]Aa is empty and that the a intersection of finitely many Aa is always non-empty. Then the Aa would form a sub-base of a filter on R which could have no adherent point, as this would belong to all the Aa, since the A^ are closed. If R is not compact, there is a filter {Fa} with no adherent points. The sets Fa form a collection of closed subsets with [\ Fa empty, al- _ a though the intersection of finitely many Fa is always non-empty. Taking complements, we immediately get (3) A space R is compact if and only if every cover of R by open sets contains a finite sub-cover. A subset M of a Hausdorff space R is said to be compact if it is a compact space in the induced topology. M is thus compact if every filter on M has at least one adherent point in M. Since the closed (respectively open) sets of M are intersections of closed (respectively open) sets of R with M, we also have (4) A subset M of a Hausdorff space R is compact if and only if a) every collection {Aa} of closed subsets of R with f](Aar\M) empty a contains a finite collection {Aa.} with [}(Aa.nM) empty, or if and only if a b) every cover of M by open subsets of R contains a finite sub-cover. From §2, 5.(3) there follows immediately (5) M is compact if and only if every net on M has an adherent point in M. 2. Properties of compact sets. (1) Every compact set is closed. If x0 is a closure point of the compact set M czR and if {Ua} is the neighbourhood filter of x0, then {UanM} is a filter on M, which can only have x0 as adherent point, since R is Hausdorff. By the compactness condition, x0 must thus lie in M. A set McK is called relatively compact if its closure M is compact. Every filter on M then has an adherent point in M cz R. (2) Every subset of a compact set M is relatively compact, and every closed subset of M is compact. (3) Every compact space R is regular. Were this not the case, there would be a point x0 in R with an open neighbourhood U, for which Aa n (R — U) would be non-empty, for every closed neighbourhood Aa of x0. By § 1, 5.(1), [\A(X={x0}, so that, a as x0$R~U, the closed sets Aan(R~U) have an empty intersection. 2 Kothe, Topological Vector Spaces 1
18 § 3. Compact spaces and sets Since every finite collection of them has a non-empty intersection, we obtain a contradiction to 1.(2). (4) The union of finitely many compact sets is compact. n This follows directly from the covering property 1.(3); if M = (J M{ i= 1 is covered by open sets, each M,- is covered by finitely many open sets, and so therefore is M. (5) The continuous image A(R) of a compact space R in a Hausdorff space is again compact, so that the mapping A is closed. Let A be a continuous mapping of a Hausdorff space R] into a Hausdorff space R2> The image A(M) of a compact (respectively relatively compact) set M^RX is again compact (respectively relatively compact). If © = {G/j} is a filter on A(M), the sets A{~l){Gp) generate a filter A{~1\<&) on M. The compactness condition implies the existence of an adherent point x0 of ,4(_1)(@>). A{x0) is then an adherent point of 05, so that A(M) is compact. (1) and (2) establish the other assertions. As a special case we obtain from § 1, 7.(2) (6) A continuous one-one mapping of a compact space onto a Hausdorff space is a homeomorphism. If R is compact under the topology % every coarser Hausdorff topology coincides with X. The second assertion follows from the first by considering the identity mapping of R onto itself. 3. Tychonoff's Theorem. This says (1) The topological product R—TIR^ of arbitrarily many compact spaces is again compact. Proof. By 1.(1) we must show that every ultrafilter <$={FP} on R converges. We form the projections ga in each R^ (cf. § 2, 6.). These are again ultrafilters. By hypothesis, each ga converges to an xae#a. By §2,6.(3), 5 converges to the element x of R whose components are the xa. In particular, by 1.8. every parallelotope S$A is compact. 4. Other concepts of compactness. The concept "compact" can be weakened in various ways. If Xa is some cardinal number, a subset M of the Hausdorff space R is called Xa-compact if every filter with a base of at most Xa sets of M has an adherent point in M. This is the case if and only if every net x(i of at most Xa elements of M always has an adherent point in M. The proof of this is given by the generalization of § 2,5. (3) to filter-bases.
5. Axioms of countability 19 The criteria 1.(2) and 1.(3) are also carried over, if the cardinal of the collection of closed sets (respectively of the collection of open sets of the cover) is at most Xa. For X0-compact, we also say countably compact. If Mis count- ably compact, every sequence of points of M has at least one adherent point. On the other hand if this is the case M is countably compact: indeed if g is a filter on M with countable base Fh i=l,2,..., then a n sequence xne f] Ft has an adherent point in M, which is also an ad- i=l herent point of all the Ft. A set M is called sequentially compact if every sequence of points of M contains a subsequence which is convergent to a point of M. A sequentially compact set is always countably compact. The converse is not true, for there even exist compact sets which are not sequentially compact, and equally there exist sequentially compact sets which are not compact. In the parallelotope ^PA, A uncountable, the set M of all x = {£a} with only countably many coordinates different from zero is sequentially compact, but is not compact, since M = ^A If A is the interval [0,1] of the real line, we can consider *PA as the set of all functions on [0,1] with values in [0,1]. If fn(x) is the function which goes linearly from 0 to 1 in every sub-interval \_k- \Q~n,(k+ 1)10~") of [0,1], the sequence fn(x) has no convergent subsequence, so that S$A is compact, but not sequentially compact. In P" all these concepts of compactness coincide. The corresponding "relative" concepts are obtained if the adherent points (respectively limits) are only required to lie in R. We shall later be concerned with the question of finding conditions which ensure that compactness or relative compactness follows from one of these weaker concepts. We mention one of the properties resulting from these ideas, which follows directly from the definitions (1) // jV1=^N2=>••• is a decreasing sequence of closed non-empty subsets of a countably compact or sequentially compact set M, then f]Nt is non-empty. i 5. Axioms of countability. In the following we shall again suppose that R is HausdorfT. First axiom of countability. Every neighbourhood filter of a point of R has a countable base. The base can then be chosen as a decreasing sequence Ux => U2 => ••*. Every subspace also satisfies the first axiom of countability.
20 § 3. Compact spaces and sets (1) If R satisfies the first axiom of countability, every closure point of a subset M of R is limit of a convergent sequence of points of M. To prove this, form the sets M n Ut from the basic neighbourhoods Ut of a closure point x0, and choose a point xt from each. A sequence convergent to x0 is obviously obtained. In such spaces countably compact sets are sequentially compact. Furthermore it follows from §1,7.(3) that sequentially continuous mappings are continuous. Second axiom ofcountability. R has a countable basis. For the definition of basis, see §1,1. Every space which satisfies the second axiom also satisfies the first. Every subspace again has a countable basis. (2) Every open cover of a Hausdorff space with a countable basis contains a countable subcover. Proof. Let R be covered by the open sets Qa and let Oh i = 1,2,..., be a basis of the open sets. Let Oin, n= 1,2,..., be the collection of those Ot from which all the Qa can be formed by taking suitable unions. For each Oi there is thus at least one Qa with Ot a Qa . But then a n n To know whether a set is compact it is therefore only necessary to examine countable covers by open sets, i.e. (3) In a Hausdorff' space with a countable basis every countably compact set is compact. In such a space the compact, sequentially compact and countably compact sets thus coincide. In these spaces, moreover, filters and nets are superfluous, and everything can be analysed using sequences of points. We remark that (3) is not always true for Hausdorff spaces satisfying the first axiom ofcountability (cf. Bourbaki [5], Vol. 4, p. 32, example 21). (4) // the Hausdorff space R has a countable basis {Ut}, every basis {Ka} has a countable subsystem which is again a basis. For every Ut is, by (2), union of countably many Vaik, fc=l,2,..., and these, taken together, also form a countable basis. 6. Locally compact spaces. A Hausdorff space is called locally compact if every point has a neighbourhood whose closure is compact. Every compact space is locally compact. P" is locally compact but not compact, and every discrete topological space is locally compact. Every closed subspace of a locally compact space is, by 2.(2), again locally compact.
6. Locally compact spaces 21 (1) Every locally compact space is regular, and the compact neighbourhoods form a base of neighbourhoods. Every point x has a compact neighbourhood U. If V is any neighbourhood, Vn U is a neighbourhood of x in the compact space U, which is regular, by 2.(3). Thus Vn U contains a closed neighbourhood W of x for the topology of U. W, being the intersection of a neighbourhood in R with U, is itself a neighbourhood of x in R. W is compact, by 2.(2), and so is closed in R. Thus the compact neighbourhoods of x form a base of all the neighbourhoods. It follows directly from (1) that every open subspace of a locally compact space is again locally compact. The topological product of finitely many locally compact and arbitrarily many compact spaces is again locally compact, by Tychonoff's theorem. If x0 is a point of a compact space R, R~x0 is clearly locally compact in the topology induced by R. Conversely (2) Alexandroff's theorem. Every space R which is locally compact and not compact can be enlarged by the addition of one point to give a compact space, the one-point compactification of R. Proof. Let R' be the space consisting of the points of R and one further point z. We define the closed sets of R' to be all the compact sets K of R together with the sets Akjz, A closed in R. The axioms (Al) and (A 2) are clearly satisfied, so that R' is a topological space. The subspace topology induced on R is the original one, since the intersections with R of the closed sets of R' which have just been defined are exactly the closed sets of R. R' is Hausdorff: it is only necessary to show that there are disjoint neighbourhoods of xgR and z. If U is a compact neighbourhood of x in R, U is closed in R', so that R' ~U is an open neighbourhood of z which has no point in common with U. R' is compact: given closed sets £a with an empty intersection, they cannot all be of the form Auz, for otherwise z would lie in their intersection. If B^o = K0cz R, then P)(K0n#J is also empty. The sets a K0nBa are closed sets of the compact set K0, however, so that finitely many of them have an empty intersection. (3) The compactijication of (2) is unique up to homeomorphism. It is enough to show that the closed sets of a compactification R = Ruz' must coincide with the sets K and Auz', K compact in R and A closed in R. R, being compact, is Hausdorff, so that z is a closed set and the sets K and Auz' are closed in R. If conversely B => z is closed, then B = (B~z')vz'. Thus B~z' has as closure points only elements of B~ z', and possibly z', and is therefore closed in R.
22 § 3. Compact spaces and sets Since every subspace of a regular space is regular, (1) is also a consequence of (2). The point z adjoined to R is called the point at infinity. A space which is locally compact but not compact is said to be countable at infinity if it is the union of countably many compact sets. (4) A space which is locally compact but not compact is countable at infinity'if and only if the point z at infinity in the compactification R' has a countable base of neighbourhoods. Proof. The condition is sufficient, for if Vn is a countable open base of neighbourhoods of z, the countable collection of compact sets R' ~ Vn cover the space R. Conversely, let KlaK2<^'" be a covering of R by countably many compact sets (by forming finite unions an increasing sequence can always be obtained). Every point of K1 has a relatively compact open neighbourhood. The union U1 of finitely many of these neighbourhoods cover K1. In the same way we find an open relatively compact set U2 which covers U1 uX2, and so on. In this way we obtain a sequence £/_„, with £/„_! cz Un, which covers R. We now show that the Vn = R' ~ Un form a base of neighbourhoods of z. From the definition of the closed sets of R' it follows that the sets R' ~K, K compact in R, form a base of neighbourhoods of z in R'. It is therefore enough to show that for every compact subset K there exists a Un with K a Un. X, being a compact set, is covered by finitely many Uk, and so by one (/„, for sufficiently large n. 7. Normal spaces. The properties of being a Hausdorff or regular topological space are not sharp enough for many purposes. A Hausdorff space is said to be normal if it satisfies the condition (N) // Ax and A2 are two disjoint closed subsets of R, there always exist two disjoint open subsets Ul => Ax and U2 => A2. An equivalent condition to (N) is (N') // A is a closed subset of R, and if U is open and U => A, then there is an open neighbourhood V of A with Kc [/. Proof. Suppose (Nr) holds. If A and B are disjoint and closed, R~B is an open neighbourhood of A, and so by (N') there exists U1 => A with U^B empty. U1 and R^Ul are disjoint open neighbourhoods of A and B. Conversely suppose that (N) holds. Applying (N) to the closed sets A and R~U, where U is an open neighbourhood of A, open sets Ux^ A and U2=> R~U are obtained, with U1 n U2 empty. But then Ulr\(R~U) is empty, so that U1 a U and (Nr) is satisfied.
2. Metric space as a topological space 23 Taking A to be one point, (N') gives (1) Every normal space is regular. Subspaces of normal spaces are not always normal, and there are locally compact spaces which are not normal. On the other hand (2) Every compact space is normal. The closed subsets of a compact space R are again compact. Let A and B be two disjoint compact subsets of R. Since R is regular, by 2.(3), for each xeA there exists an open neighbourhood U(x) with U(x)nB empty. As x runs through the whole of A, the U(x) form an open cover n n of A, and, by 1.(4), A a (J U{xt). A' = (J U(xt) is a closed set disjoint i = i i = i n from B. Thus L^ = (J U(xt) and [/2 = #~;4' are disjoint open sets i= 1 with Ul^ A and £/2 => #• § 4. Metric spaces 1. Definition. A set R is called a metric space if a real number \x,y\, the distance between x and y, is defined for every pair x,y of elements of R, with the following properties: (Dl) \x,y\^0, (D2) |x,j/| = 0 if and only if x = y, (D3) \x,y\=\y,x\, (D4) |x,z | ^ |x,}/1 + \y, z| (triangle inequality). We also say that a metric is defined on R by the function \x,y\. Every subspace of a metric space is again a metric space, using the same definition of distance. A one-one mapping x-»x' of a metric space R onto a metric space R' is called isometric if \x,y\ = \x\y'\ holds for all pairs x,y. The simplest example is P" with distance defined by |x,j>|= 1/ X toi~£;l2- This is called ^-dimensional Euclidean space. ', = 1 00 The set of all real sequences * = (£i,£2>--) with £|£t|2<oo forms a metric space with distance defined by \x,y\ = / Z to/-£fl2- This space is called Hilbert space. r , = 1 2. Metric space as a topological space. If x0 is a point of the metric space R, the set of all yeR with |x0,j/|<r (respectively \x0,y\^r) is
24 § 4. Metric spaces called the open (respectively closed) ball of radius r and centre x0. The set of y with \x0,y\ = r is called the sphere of radius r about x0. If A, B are two subsets of R the greatest lower bound inf \x,y\ is xeA,yeB called the distance between the two sets. The number sup \x,y\ is x,yeA called the diameter of the set A. A set A is said to be bounded (with respect to the metric) if its diameter is finite. (1) Every metric space becomes a Hausdorjf topological space when the open balls with centre x are taken as a base of neighbourhoods of the point x. Every metric space is a normal space with a countable base of neighbourhoods at each point. Proof. Axioms (HI) and (H2) of §1.3 are clearly satisfied. (H3) results from the following: if y lies in the open ball Kr(x) of radius r about x, then d=\x,y\<r, so that there exists e>0 with d + e<r. But then KE(y) lies inside Kr(x), since if zeKE(y), |x,z|:g \x,y\ + \y,z\<d + e<r. Further if x=ty, d=\x,y\=£0. The open balls Kd/2(x) and Kdl2(y) have no common point z, for otherwise |x,j/|:g |x,z| + \z,y\ would be less than d. Thus (H4) is satisfied. Axiom (N) for a normal space is satisfied: if A and B are disjoint closed sets, the set U of all x with |x,y4|<-j|x,£| and the set V of all y with |j/,#|<||j;,,4| are disjoint open neighbourhoods of A and B. Finally the balls about x with radii 1 /n, n = 1,2,..., form a countable base of neighbourhoods of x. Two isometric spaces are homeomorphic, but the converse does not always hold. 3. Continuity in metric spaces. Since a metric space satisfies the first axiom of countability, the results of § 3, 5 apply. A mapping A from a metric space R1 into a metric space R2 is therefore continuous if x(n)->x0, i.e. |x(n),xo|-»0, always implies that Ax{n)^>Ax0. This can clearly also be expressed in terms of c and 5: A is continuous if for each £>0 there exists d>0 for which |x,x0|<<5 always implies that \Ax,Ax0\<8. Closure points and accumulation points of a set are always limits of convergent sequences of elements of the set, and so a limit of limits of a set is again limit of a sequence of elements of the set. The metric allows what has not been possible in the topological spaces considered up to now, namely the comparison of the size of neighbourhoods of different points. This produces a series of further concepts. For example, uniform continuity of a mapping A can be defined as follows: A is uniformly continuous if for every e>0 there exists (5>0 such that \Ax,Ay\<s whenever |x,j/|<(5.
4. Completion of a metric space 25 The concept of a Cauchy sequence can also be defined: x(n\ rc=l,2,..., is a Cauchy sequence if for each s>0 there exists n0(s) such that \xim\x(n)\<s whenever m,n = n0. A metric space is called complete if every Cauchy sequence has a limit in R. 4. Completion of a metric space. Following the method of establishing the real numbers of Cantor and Meray, we prove the following theorem (1) Every metric space can be embedded in a smallest complete space R, which is unique up to isometry, and is called the completion ofR. For every two Cauchy sequences x = (x{n)) and y = (y(n)),lim\x(n\y(n)\ = |x,j/| always exists, for since \x{m\y{m)\ = \x{m\x(n)\ + \x(n\y(n)\ + \y(n\y(m)\ we have ||x(m))};(m)| _ |x<«>j},<«>|| = |x(«)jX(»)| + \yln\yM\. If |x,x| = 0 and |y,j>| = 0, then |x,j/| = \x,y\: for from \xin\y(n)\ = \x(n\x{n)\ + \x{n\y(n)\ + \y(n\y{n)\ lim\xin\yin)\^lim\xin\y(n)\ follows directly, and the interchange of x, y with x, y establishes the assertion. The relation |x,}/| = 0 defines an equivalence relation on the set of all Cauchy sequences. We denote the class to which x belongs by x. By what we have just shown, a unique real number \x,y\ = lim |x(w),y(w)| ^ n-* oo is associated with each pair of classes. Under this the set R of all classes becomes a metric space: (Dl), (D2) and (D3) are trivially satisfied, and (D4) follows from \x(n\zin)\ = \x{n\y(n)\ + \y(n\zin)l by taking limits. If we assign the element aeR to the class a determined by the Cauchy sequence (x{n)) with x(n) = a, an isometry is obtained, and we can identify R with a subspace of R. R is dense in R, for if x is the Cauchy sequence (x(n)), 3c = limx("): indeed \x,x(n)\ = lim \x{m\x{n)\^c for n^no(r,). R is complete: let (xn) be a Cauchy sequence in R. For each xn there is an a{n)eR with \xn9ain)\<l/n. Since |a<->,a<">| ^ |a<™>,xm| + |xm,xj + |x„,a<">| the a{n) form a Cauchy sequence a in R, and from \a,xn\ g \a,a{n)\ + \a(n\xn\ it follows that a = limx„. Since R is dense in R and the distances \x,y\ are uniquely defined as lim|x(n),};(,1)|, they are the same in each extension of R, and R is uniquely determined up to isometry as the completion of R.
26 § 4. Metric spaces 5. Separable and compact metric spaces. A topological space is called separable if there is a countable set of points which is dense in it. (1) A metric space R is separable if and only if it has a countable basis. In one direction the result holds for general topological spaces: if Oh i= 1,2,..., is a countable basis and xtEOi9 the x{ are clearly dense in R. If conversely xt is a sequence which is dense in the metric space R, we shall show that the open balls Kp(xt) with rational p form a basis: it is enough to show that each xeR has a base of neighbourhoods formed of suitable K()(xt). But given Kp(x) and |x,xf|<p/2, Kp/2(xt) is a neighbourhood of x contained in Kp(x). A compact metric space is called a compactum. Every compactum is complete, as § 3,1.(3) shows immediately. (2) Every compactum, and equally every relatively compact subset M of a metric space, is separable and has a finite diameter. For every n there are finitely many open balls K1/„(x/")), i= l,...,N(w), which cover the whole of M. The second part of the proposition follows from this. Further the countably many xf] taken together clearly form a dense set in M. (3) Every sequentially compact metric space is a compactum. By § 3, 5.(3) it is enough to show that R has a countable basis, and thus that it is separable. For each n there exist finitely many xj.n) with the property that each x is at a distance of less than \/n from a suitable x{"\ For otherwise there would be for some n0>0 a sequence x(n) with \x("\x{m)\ ^ l/nQ for all n, m, and this sequence would have no convergent subsequence. Again the countable set of all x\n) is dense in R. Applying § 3, 5.(3) again, we see that (4) In a metric space the countably compact, sequentially compact and compact sets coincide. A metric space R is called precompact if its completion R is compact. This is the case if and only if every infinite subset of R contains a Cauchy sequence; it is easy to show that R has this property whenever R does, and thus is sequentially compact. A subset M of a metric space is called totally bounded if for every s>0 there is a finite cover of M by sets of diameter ^s. (5) A subset of a metric space is precompact if and only if it is totally bounded. Proof. Every precompact set is totally bounded, since M, and so also M, is covered by finitely many sets Nt of diameter <^e. If conversely M is totally bounded and infinite it is covered by finitely many sets of diameter ^ 1, and so in one of these sets there
6. Baire's theorem 27 exist infinitely many xj^eM with Ixj^xj^lrgl for all i, /c. From these in turn we can find infinitely many x\2) with |xj-2),x[2)| ^?, and so on. The diagonal sequence x[l\x{2\... clearly forms a Cauchy sequence. (6) // Nx => N2 => • • • is a decreasing sequence of closed sets in a 00 complete metric space, and if their diameters tend to zero, then [\ Nt consists of one point. i = 1 For every sequence xfeNf is a Cauchy sequence, and all such sequences converge to one and the same point. 6. Baire's theorem. We give this in several versions. 00 (1) // the complete metric space R is the union [j Mt of countably i= 1 many subsets, then at least one Mt contains an entire ball Were this not the case, each Mf would be nowhere dense in R (cf. § 1, 5), and there would then be a closed ball Kt of diameter ^1 with Kx nMx empty, a closed ball K2 in Kt of diameter ^{ with K2c\M2 empty, and so on. The decreasing sequence Kx dK2d- would by 5.(6) have one common point, which could not belong to any Mf and so could not belong to R. This theorem of Baire's also holds for locally compact spaces. 00 (2) // the locally compact space R = [J Mh then at least one Mt contains a non-empty open set. i=l The proof proceeds analogously: were all the Mt nowhere dense in R, then there would be a non-empty open set Ox with Ox nMx empty. By §3,6.(1) R is regular, so that Ox would contain a closed subset Kx with interior points and Kx nM1 would be empty. We could take K{ to be compact. In Kt there would be a compact subset K2 with K2 n M2 empty, and so on. By § 3,4.(1) the sequence KjD^d- would have 00 a non-empty intersection which again would not lie in R = [j Mf. i=i (1) and (2) can be extended to (3) // an open subset 0 of a complete metric or locally compact space R is the union of countably many subsets Mh then at least one Mt n 0 contains a non-empty open subset of R. If R is locally compact, 0 is also locally compact in the induced topology, by § 3,6. Hence the assertion follows directly from (2). Suppose that R is a complete metric space. Since R is regular the 00 open set 0 contains a closed ball K. Then K= [j Nh where Nt = MinK. Applying (1), we find an Nt which contains a closed ball, and a fortiori an open ball K0. Consequently K0a MtnK a M. n 0.
28 § 4. Metric spaces If the M f are all nowhere dense in R, there is no non-empty open set 00 contained in (J Mh so that (4) In a complete metric or locally compact space R the complement of a countable union of nowhere dense sets is dense in R. A countable union of nowhere dense subsets of R is also called a set of the first category (Baire) or meagre (Bourbaki). If a set is not of the first category in R, it is said to be of the second category. In this case Baire's theorem takes the following form: (5) Every open subset of a complete metric or locally compact space R is of the second category in R. 7. The topological product of metric spaces. If Rl^..,Rn are finitely many metric spaces, we define, for elements x = (xx,...,x„), y = (yi,...,y„) of RlX"xRn. n \x,y\= Z tattl; i= 1 if Rl9R29... is a sequence of metric spaces, we define in a corresponding 00 way, for x,ye TT Ri9 i=l2l 1+ 1x^1 In either case it is easily confirmed that in this way a metric space is produced. Instead of 1/21 one could take any cf>0 for which 00 £ ct< oo. To be sure, by doing this, different metrics are given, but we i= 1 have 00 (1) Every such metric defines the product topology on TT Rt. Proof. An elementary argument shows that metric convergence x(")_>x(0) jn y\ ^ js eqUivaient to the metric convergence xi-n)->x[0) of the components in each Rt. By §2, 6.(1), convergence in the topological product sense also has this property, so that the two concepts of convergence are identical. Since the first axiom of countability holds in the Rh it also holds in the topological product, and thus, by § 3, 5.(1), every closure point is limit of a convergent sequence. It follows that the closed sets of R in the metric sense and in the product topology sense coincide, and so therefore do the topologies. 00 (2) TT Rt is complete if and only if the metric spaces R( are complete.
1. Definition 29 The parallelotope ty™ (cf. §1,8.) is also called the Hilbert par- 00 allelotope, since it is homeomorphic with the product R= T\ Rh where Rt is the interval [0,1//], for which we have (3) R is a closed subspace of Hilbert space, and its topology is induced by the metric on Hilbert space. Proof. The first part of the assertion is trivial. If \xjn)\ = \ J|^._^»)|2-*0, ioxx,y{n)eR, then |^-^n)|->0 for each L 00 Conversely, if this holds, then given £>0, £ |^f-^")|2<82/2 for a i = m m- 1 sufficiently large m and for each n, and £ \<t{ — rj^l2<s2/2 for suffi- /= i ciently large n, so that \x,y{n)\<e. The assertion now follows from the fact that the two concepts of convergence agree, as in the proof of (1). R is called the Hilbert cube. § 5. Uniform spaces 1. Definition. The topological spaces in which we shall be interested later are by no means all metric spaces, but ideas can be used for them which are similar to those used for metric spaces in the preceding paragraph. These spaces, which, like metric spaces, have an additional structure defining the topology, were first considered by A. Weil [1]. Let Kbea set, and RxR the set of ordered pairs (x,y) of elements of R. If N is a subset of RxR, with elements (x,y), we denote the set of all transposed pairs (y,x) by N'1. The product MN of two such sets consists of all (x,z) for which there exists a y with (x,y) in M and (y,z) in N. NN is also written as N2. Using this terminology, we make the following definition: R is a uniform space when a filter 91 is given on RxR with the following properties: (VI) Every NeW contains the diagonal, that is, the set of all (x,x) with xeR, (V2) // NeM, so does AT1, (V3) For each Ne$l there exists MesJl with M2 a N. The sets of the filter are called the vicinities of the uniform space; if (x,y)eN, x and y are called close, of order N. $1 defines a uniform structure or uniformity on R.
30 § 5. Uniform spaces A base of the filter 91 is called a base of the uniform space or of the uniformity. A filter base © of subsets of R x R forms the base of a uniformity if (V 1) und (V 3) are satisfied for © instead of % together with (V2') // JVgS, there exists iV'e© with N' c AT1. There is always a base of symmetric vicinities, i.e. those with N = N~l9 since the symmetric vicinity MnM~l is contained in the vicinity M. Every metric space is a uniform space, taking as base the countably many sets Nl/n of all (x,y) with \x,y\<l/n. Every subspace S of a uniform space R becomes a uniform space when the intersections with S x S of the sets of a base for R are taken as base. This is the uniformity induced on S by R. Two uniform spaces R and R' are called isomorphic or equivalent if there is a one-one mapping of R onto R\ and so of R x R onto R' x /?', under which the filters defining the uniformities are transformed into each other. Two metric spaces can be equivalent without being isometric. For example, one need only multiply all the distances in a metric space by a fixed positive factor to obtain a new metric equivalent to the original one. If two uniformities on a set R are given by filters 91 x and$ft2> the uniformity given by 9^ is called finer (respectively coarser) than that given by$ft2> if 911 => 9t2 (respectively 9t2 =>5Ri). 2. The topology of a uniform space. If x is a point of the uniform space R and if N is a vicinity, the set UN(x) of all y with (x,y)eN is defined to be a neighbourhood of x. The neighbourhoods defined in this way for all Ne9t and for all xeR satisfy axioms (Nl) to (N4) of § 1,2: (Nl) follows from (VI), and (N2) and (N3) follow from the filter properties of 91. To show (V4) we choose an M2 c N9 using (V3). Then if yeUM(x0)9 every point z of UM(y) is in UN(x0)9 since from (x0,y)eM and Q/,z)eM it follows that (%z)eM2ciV. Thus C/N(x0) is also a neighbourhood of y. The topology defined in this way is called the topology of the uniform space R. In the case of a metric space we obtain the topology introduced in § 4, 2. once again. A topology is defined on R x R as well, namely the product topology. (1) Every uniformity has a base of symmetric vicinities which are closed in the topology of RxR. If N is a vicinity and M is a symmetric vicinity with M3 <= N9 then M aN: if (x0,j;0) is a closure point of M there is an (x9y)eM with (x,x0)eM and (j/,j/0)eM, so that (x0,j/0)eM3 <= Af.
3. Uniform continuity 31 (2) The topology of a uniform space is Hausdorff if and only if the further axiom (V4) The intersection of all Ne$1 is the diagonal, is satisfied. If the intersection is the diagonal, the diagonal is, by (1), also the intersection of all the closed vicinities, and is itself closed. If x^y, there is a neighbourhood U x V of (x,y) which has no point in common with the diagonal, so that the neighbourhood U of x and the neighbourhood V of y have no common point. On the other hand if R is Hausdorff and x=Ny, then there is a UN(x) which does not contain y. This means that (x,y) does not lie in N, and so the intersection of all vicinities is the diagonal. If the diagonal is itself a vicinity, the uniformity is called the discrete uniformity; we obtain the discrete topology on R from it. If we take the sequence l/«, w=l,2,..., as a subset S of the metric space P, the induced uniformity on S is not the discrete one, although S is discrete as a topological space. If M is a closed vicinity of the Hausdorff uniform space R, the set of all (x0,y) in M is closed in RxR, being the intersection of the closed sets {x0} x R and M, so that U^(xQ) is closed in R; (1) gives (3) Every Hausdorff uniform space is regular. (4) // A is any subset of Rx R, A = f]N AN, as N runs through all the symmetric vicinities ofR. N If Nt and N2 are two symmetric vicinities and if N = N{nN2, then the neighbourhood UNl(x)x UN2(y) of (x,y) contains the neighbourhood UN(x)x UN(y); hence these neighbourhoods form a base of neighbourhoods of (x,y), as N varies. If (x0,y0) is a closure point of A, there is, for each N, an (x,y)eA with (x,y)eUN(x0) x UN(y0), and by the symmetry of N we also have that (x0,y0)eUN(x) x UN(y), so that (xo>yo)ENAN and Aaf)NAN. If conversely (x0,y0)ef)NAN, N N there is, for each N, an (x,y)eA with (x0,yQ)eUN(x) x UN(y), so that (x,y)eUN(x0)x UN{y0), i.e. (x0,y0)eA. Corresponding to (1) we have (5) Every uniformity has a base of symmetric vicinities which are open in the topology of RxR. It is enough to show that the interior of a vicinity contains a vicinity. If M is a symmetric vicinity with M3 <= JV, and if (x0j/0)eM, then all the points (x,y) of the neighbourhood UM(x0) x UM(yQ) lie in N, so that (xQ,y0) is an interior point of N. 3. Uniform continuity. As can be seen in the proofs which have just been given, a set whose pairs of points belong to a sufficiently small
32 § 5. Uniform spaces symmetric vicinity corresponds to a set of diameter <e in a metric space; such a set is called small oforder N. A symmetric vicinity M with M2 <= Af then corresponds to a diameter <a/2. A mapping Ax — x' of a uniform space R into a uniform space R' is called uniformly continuous, if to every vicinity N' in R' there corresponds a vicinity N in R with the property that (x',j/)eAT whenever x,j/ are two points with (x,y)eN. It is clear that every uniformly continuous mapping is continuous. Every isomorphism of two uniform spaces is uniformly continuous in both directions. If real functions fp(x),peB are defined on a set R, they can be used to induce a uniformity on R. For every finitely many fPi, i= l,...,rc, and £>0, we form the set N of all pairs (x9y)eRxRmth\fPt(x) — fp.{y)\<e for i—\9...9n. These sets N form the base of a filter of vicinities on RxR. The fp are clearly uniformly continuous with respect to this uniformity on R, and it is the coarsest uniformity on R with this property. The uniformity is Hausdorff if and only if every two points x9y of R are separated by at least one fp, i.e. fp(x)+fp(y). This idea can be generalized directly to the case where a collection of maps Ap from R into uniform spaces Sp is given. The sets of all (x9y) with (Ap.(x)9Ap.(y))eNPi>y. for i=l,...,n, where NPiyy. is an arbitrary vicinity in Sp., are then taken as vicinities. 4. Cauchy filters. Cauchy filters and Cauchy nets in uniform spaces correspond to Cauchy sequences in metric spaces: a filter {FJ on R is called a Cauchy filter if to each vicinity N there corresponds at least one Fa which is small of order N9 i.e. (x9y)eN for all x,j/eFa. An adherent point of a Cauchy filter is a limit of the filter, and in a Hausdorff uniform space a Cauchy filter has at most one adherent point. Every convergent filter is a Cauchy filter. If every Cauchy filter on R is convergent, R is said to be complete. In an analogous way, a net xa9 aeA is called a Cauchy net if to each vicinity N there corresponds a j5eA for which (xy,xd)eN whenever From the connection between nets and filters (§ 2, 4. and § 2, 5.) there follows directly (1) R is complete if and only if every Cauchy net has a limit. For metric spaces we have (2) A complete metric space is complete as a uniform space. If {Fa} is a Cauchy filter on R, there is a decreasing sequence Fa„ with \x,y\^\/n for all x,yeFan. A sequence xneFan is then a Cauchy sequence in R. Its limit x0 is an adherent point of each Fa, since FanFan is non-empty.
5. The completion of a Hausdorff uniform space 33 In general a uniform space can very well contain limits of all the Cauchy sequences in it, without being complete, for the limits of certain Cauchy filters can be missing (cf. the example tyA in § 3, 4.). (3) Every discrete uniform space is complete. Every uniformly continuous mapping sends every Cauchy filter into a Cauchy filter, and equally sends every Cauchy net into a Cauchy net. Theorem § 2, 8.(3) on the extension of continuous mappings leads to (4) A uniformly continuous mapping from a dense subspace S of a uniform space R into a complete Hausdorff uniform space R' can be extended in a unique way to a uniformly continuous mapping from the whole of R. The condition of § 2, 8.(3) is satisfied. For every point xeR, the sets [/anS form a Cauchy filter g on5; its image A (ft) is a Cauchy filter on R\ which defines the image Ax. Moreover R' is regular, by 2.(3). Thus A is continuously defined on the whole of R. A is uniformly continuous on R: let N' be a closed vicinity in R\ and let N be a vicinity in R with the property that (Ax,A,y)eN' whenever (x,y)eNn(5 x 5). Let M be a symmetric vicinity with M3 <= N. If x0,j/0 are points of R with (x0,j/0)eM, then (x0,j/0) is a closure point in RxR of 5x5, and in particular of the set of all (x,j/)in5x5 with (x,x0)eM and (j/,j/0)eM. If (x,j/) belongs to this set, (x,j/)eM3ciV, so that (Ax,Ay)eN'. Since (Ax0,Ay0) is a closure point of the (Ax,Ay), and since N' is closed, (Ax0,Ay0) lies in AT. Thus (AxQ,AyQ) belongs to N' for all (x0,j/0)inM. 5. The completion of a Hausdorff uniform space. Let 93 = {Afa}, aeA, be a base of symmetric vicinities for the uniformity on R [cf. 2.(1)]. If we put a ^ j8 when Na => Np, A becomes a directed set. 4.(1) can be sharpened to give (1) The uniform space R is complete if and only if every net with index set A has a limit in R. It is enough to show that every Cauchy filter g = {Fa} has a limit in R. For every aeA we choose an Fae ft which is small of order Na, and an xaeFa. If /J^a, y^a and zeFpnFy, then (xp,z) lies in NpczNx and (xy,z) lies in Ny <= Na, so that (xp,xy) lies in N2a. The net (xj is thus a Cauchy net and by hypothesis has a limit x0 in R. Now let Nde%> be given and let Npe93, with N* <=: Ns. There exists an xff with <r^p and (x0,xJeiVp. If zeFac\Fp, (xa,z)eNp. If yeFp, then (z,y)eNp. Hence (x0,y)eNp cz A^, for all yeFp. But this means that $ has x0 as limit. As analogue of § 4, 4. we now show (2) Every Hausdorff uniform space R may be embedded in a smallest complete Hausdorff uniform space R, which is unique up to isomorphism. R is called the completion of R. 3 Kothe, Topological Vector Spaces I
34 § 5. Uniform spaces Proof, a) Definition of R. We consider the Cauchy nets (xj, aeA, on R, where A is the index set of the base fixed above. Two such nets (xj and (j/a) are said to be equivalent if for each symmetric vicinity N of the uniformity on R there exists an index (}(N) for which (xy,yy)eN for all 7 ^ /}. If (xj is equivalent to (j/a), and (j/J to (zj, then (xJ is also equivalent to (zj: for if M is a symmetric vicinity with M2 <= JV, it follows from the fact that (xy,yy)eM and (yy,zy)eM for sufficiently large 7 that (xv,zy)eM2 <= N. We collect equivalent Cauchy nets together into equivalence classes, which we define to be the points of R. The class in which the Cauchy net (xj lies will be denoted by x. b) The uniform structure of R. If M is a symmetric vicinity of R, let M denote the set of all pairs (x,y) with the property that for each symmetric vicinity N there exists an index /?(N) such that (xy9yy)eNMN, for y^.ft(N). This definition is independent of the representatives of the classes: suppose that (zj is equivalent to (j/J. If Nj <= N, (xy,yy)eNlM Nx for 7^j8(N1). If, further, 7 is chosen large enough for (yy,zy) to belong to Nu then (^zJe^MAfJcNMiV. Thus (x,z) belongs to M. The M form the base of a uniformity on R: (V1) is clearly satisfied, and so is (V 2), as M~l = M, because of the symmetry of M. M c A/\ n N2 implies that M c NlnN2, so that the M form a filter base on RxR. Finally, if M3 <= JV, it is easily confirmed that M2 c N, so that (V3) also holds. c) RisHausdorff.li (x,y) lies in every M, (xy,yy)eM3 for 7^jS(M), for all M; (xj and (yj are thus equivalent. The intersection of all the M is therefore the diagonal. d) R is a subspace of R. We denote the equivalence class containing the Cauchy net xa = a, aeR, by a. (a,b) lies in M if and only if (a9b)ef)NMN. By 2.(4) this means that (a,b)eM. If we identify a and N a, then, from what we have just shown, Mn(RxR) = M, so that by 2.(1) the uniformity induced on R by R coincides with the original structure. e) R is dense in R. If x is a general element of R and (xj is a net defining x, there is for each M a j8(M) for which (x7 ,xy)eM for all y\Y^P- But this means that (x,xy)eM, when xy is considered as an element of R. f) R is complete. If 93 = {Afa} is the base defined at the beginning, then, by b), © = {Na} is a base for the uniformity on R. By (1) it is enough to show that every Cauchy net (xj, aeA, has a limit in R. By e) we choose for each xa a point za in R with (xa,za)eiVa. (zj is then a Cauchy net in R: let peA be specified so that Np3 <= Na. Then (jcy,xy')GNp whenever 7,7'^sup(j5(p), p). Since, moreover, (xy,zy)eNp and (xy>,zy.)eNp, it follows that (zy,zy,)eNp c Na, i.e. (zj is a Cauchy net z. Since (z,zy)eNp
6. Compact uniform spaces 35 for sufficiently large y, and since (zy,xy)eNp and (xy,xy,)eNp, it follows that (z,xy,)eN0L for all sufficiently large /, i.e. the Cauchy net (xj has z as limit. R is thus a completion of R. Finally, the fact that R is uniquely determined up to isomorphism is a direct consequence of (3) // Rl and R2 are isomorphic uniform spaces which are dense in the complete Hausdorff uniform spaces Sl and S2 respectively, then any isomorphism between Rl and R2 can be extended uniquely to an isomorphism between Sv and S2. Let A(R1) = R2, B(R2) = Rl be the mappings given by the isomorphism. Since an isomorphism is uniformly continuous in both directions, the existence of two uniformly continuous mappings A and B which map Sl into S2 and S2 into Sj respectively follows, by extension, from 4.(4). The class of equivalent Cauchy nets in Rt corresponds to the class of equivalent Cauchy nets in R2 under the isomorphism of Rt and R2, so that A and B are uniformly continuous mappings of Sj onto S2 and S2 onto Sl respectively, and are each the inverse of the other. Consequently the vicinity filters of Sl and S2 are also sent into each other, and A and B define an isomorphism of St and S2. (4) // M is a vicinity in RxR, the vicinity M in RxR is the closure of M in RxR. The closure in RxR of the vicinities of a base in RxR form a base in RxR. If (x(0),y{0)) is a closure point of M, then for each symmetric vicinity N there exists an (xj)eM for which (x(0),x)eN and (y,yi0))eN, so that, for sufficiently large y, (x{y°\xy)eN39 (xy,yy)eNMN and (jvj^gN3; thus (xy°\y^)eN4MN4. But this means that (x(0),y(0))eM, and that M is closed. On the other hand if (x,y)eM, then for each symmetric vicinity N there exists elements xvyy in R such that (x9xy)eN, (y,yy)eN and (xy9yy)eNMN. Thus there exist zjeR with (z,t)eM, (z,xz)eiV and (t9y7)eN9 so that (z9t)eUff2(x)x V^(t)9 and M is dense in M. For the proof of the second assertion we observe that, according to the definition of the uniform structure on R, the M with symmetric M form a base in RxR. Every M contains a vicinity N of the given base in RxR, so that MdN as well, and the N form a base in RxR. 6. Compact uniform spaces. (1) In a compact uniform space every ultra- filter is a Cauchy filter. If M is a vicinity of R, R is covered by finitely many open sets M,, i=\,...,n, which are small of order M. By §2,7.(2) one of the M,- belongs to the ultrafilter, which thus contains arbitrarily small sets. 3*
36 § 5. Uniform spaces (1) also follows directly from §3, 1.(1). Every compact uniform space is complete, since every Cauchy filter has an adherent point, and thus has a limit. As in §4,5. we call a Hausdorff uniform space R precompact if its completion is compact. A Hausdorff uniform space is said to be totally bounded if for every vicinity N of R there exists a covering of R by finitely many sets which are all small of order N. Analogous to § 4, 5. we have (2) a) A Hausdorff uniform space R is precompact if and only if it is totally bounded. b) A Hausdorff uniform space is precompact if and only if every ultrafilter is a Cauchy filter. Proof, a) If R is precompact, the completion R, and thus also R, is covered by finitely many sets Mt which are small of order N. R is covered by the sets MtnR, which are small of order N since N is obtained from the closed vicinity N in R x R by forming its closure in RxR. If conversely R is totally bounded and finitely many Mf of order N cover the whole of R, then the closures Mt in R cover the whole of R and are small of order N. By 5.(4) the N form a base for the uniformity of R. If now 5 is an ultrafilter on R, then, as in (1), by § 2, 7.(2) at least one Mt is in J$f- 5 tnus contains arbitrarily small sets and is therefore a Cauchy filter, which converges in R, since R is complete. R is therefore compact. b) A precompact space is totally bounded, by a), and it follows from this, as in (1), that every ultrafilter is a Cauchy filter. If R is not precompact, R is not compact, and so there is a filter g={Fa} on R which has no adherent points. For every F* and every symmetric vicinity N we form the set F| of all zeR with the property that (x,z)eN for some xeF*. The sets F| form a filter base on R, and, as F$nR is always non-empty, the sets F%nR form the base of a filter g' on R. g' has no adherent points in R, since such a point would also be an adherent point of 5- Any ultrafilter on R which is finer than 5' is therefore not a Cauchy filter. (3) A Hausdorff uniform space R is precompact if and only if every sequence in R has an adherent point in R. The condition is clearly necessary. If conversely R is not precompact, there is, by (2) a), a vicinity N for which there is no finite cover of R by sets of order smaller than N. Let Nl be a symmetric vicinity with Ni cz N. We choose an xl and form the neighbourhood Uj^^x^. This is of order Nf. Now suppose that points xu...,xk have already been
7. The product of uniform spaces 37 chosen so that the UNl(xt) are pairwise disjoint. Were there no xk+1 for which UNl(xk + l) was disjoint from the UNl{Xi) already determined, then all the points of R would be contained in the sets UN2(xt)9 i = l,...,fc, so that k sets of order smaller than N* cz N would cover the whole of R. For the sequence xk9 k = 1,2,...,determined in this way, (xk9xm)$Nl for all fc, m, so that the sequence has no adherent point in R. (4) A complete Hausdorff uniform space is compact if it is count ably compact. For by (3) a countably compact Hausdorff uniform space is pre- compact, and compactness follows from completeness. (5) In a compact uniform space, the filter 91 of vicinities consists of all the neighbourhoods of the diagonal in RxR. Every vicinity N is a neighbourhood of the diagonal in RxR. Were there, on the other hand, an open neighbourhood N0 of the diagonal which did not belong to9l, then, for each Ne9l9(Rx R~N0) n N would be non-empty, since Nc£N0; the sets (RxR~N0)nN would thus define a filter 91' => 9t, which, because of the compactness of Rx R, would have an adherent point belonging to RxR~N09 and so not belonging to the diagonal. But 9t, and also 91', have only points of the diagonal as adherent points, since R is Hausdorff and regular. (6) Every continuous mapping A of a compact uniform space R into a uniform space R' is uniformly continuous. If A is continuous, by §2,6.(2) the mapping (x9y)^(Ax9Ay) from RxR into R' x R' is continuous. If, therefore, AT is an open vicinity in R' [cf.2.(5)], the set of the inverse images (x9y) of all (Ax9Ay)eNf is a neighbourhood of the diagonal in RxR, and so, by the preceding theorem, is a vicinity in R. 7. The product of uniform spaces. If Ra9 aeA, are uniform spaces with vicinity filters 91a, R = T\Ra becomes the uniform product of a the R^ when one takes as basic vicinities N in RxR = T\ (Ra x RJ all a sets TTMa, where Ma = RaxRa for all but finitely many a and M0L = N0Lx Na for the rest, where Na is an arbitrary vicinity in 9ta. It is easy to see that the topology defined by this uniformity on R is the product topology (§ 1, 8.). We can add the following easily established results to those already obtained in §2,6.: (1) A filter 3 on R is a Cauchy filter if and only if all its projections 5a are Cauchy filters. (2) The uniform product R = T\ Ra of the uniform spaces R^ is complete if and only if all the Rx are complete. The completion of R is the uniform product of the completions of the Ra.
38 § 6. Real functions on topological spaces This follows from (1) and § 2, 6.(3). A mapping A of a uniform space into the uniform product TT Ka is uniformly continuous if and only if all the projections A0L = P0LA are. § 6. Real functions on topological spaces 1. Upper and lower limits. If we add + oo and — oo to the set P of real numbers, we obtain a set which we denote by P. The order relations and algebraic operations in P are defined in the usual way. P becomes a topological space when we take the intervals [— oo,a], [b9 + oo] as a sub-basis for the collection $X of closed sets. The original topology of P is induced by this. The mapping x->tgx maps [ — n/2,n/2] homeo- morphically onto P, so that P is compact. The least upper bound supa and the greatest lower bound aeA _ inf a are defined for any subset A of P. They can of course be + oo aeA or—oo. If M is any set and if an element f(x) in P is ascribed to each xeM, we speak of a real function on M. If the values lie only in P, f(x) is called a finite-valued real function. Least upper bounds and greatest lower bounds of the values of the function on M are denoted by sup/(x) and inf/(x) respectively. xeM xeM Since we allow both + oo and -x as values of functions, the sum f{x) + g{x) need not be defined for all x (+ oc and — oo cannot be added). But if f{x) + g(x) is defined for at least one xeM, then the following hold, provided that the right-hand sides are meaningful: (1 a) sup (f(x) + g(xj) ^ sup/(x) + sup#(x), xeM xeM xeM (lb) inf (f(x) + g(x))Z inf f(x)+ inf g(x). xeM xeM xeM The formulation and proof of the corresponding formulae for multiplication and division are left to the reader. (2) If f(x)^g(x) for all xeM, then supf(x)^supg(x) and inf/(x) ^infg(x). *eM *eM xeM xeM A net £a on P is called monotonic increasing (monotonic decreasing) if a</? implies that £a^£p {£a^£p). (3) Every monotonic increasing (decreasing) net £a on P has a limit, and lim£a = sup£a( = inf£a). For if sup^ = y, then for each 5<y there exists a ft for which 8<€fi'^ 7 for aU /*'^ ]8, i- e- 7 is tne limit °f tne set ^a-
1. Upper and lower limits 39 If £a is an arbitrary net on P, then the net rja = sup £p derived from it is monotonic decreasing. By (3), \imrja = mf sup<^ exists; we call it a P^a the upper limit or limes superior of £a and denote it by lim £a. Similarly the lower limit or limes inferior is defined as lim £a = sup inf £«. a P^ol The inequality lim £a ^ lim £a always holds. If g = {Fa} is a filter on P, whose indices are ordered by the rule that a:g/? if Fa=> Fp, then we have the analogous definitions Hmg = limFa = lim sup £ = inf sup £, limg = limFa = lim inf £ = sup inf i. a £eF« a $eF„ We observe that in order to form lim 5 it is only necessary to use the Fy from a base for g, since the net rjy = sup £ is cofinal with the net of all rja, and so has the same limit. *eFy (4) T/z^ upper /imit o/ a net or filter on P is the greatest adherent point of the net or filter. We shall only show this for nets. Let /7 = lim^a. Since ?7 = inf?7a, and since rja is monotonic descreasing, given s>0 there is an a0 for which \rj — rja\<e/2 for a^a0. For each rja there is a <^a, with /5a^a, for which |?ya—^a|<e/2. The net ^a, a^a0, which is cofinal with £a, therefore lies in the ^-neighbourhood of rj, and ^ is an adherent point On the other hand if £ is an adherent point, £^??a for each a, so that £:ginf?7a = ?7, by (2). Now let f(x) be a real function on the topological space R. If xa is a net convergent to a then the values f(xa) of the function form a net whose upper and lower limits we denote by lim f(x) and lim f(x) respectively. x""a x""a On the other hand if U={U0L} is the filter of neighbourhoods of a, the image filter f(VL) is a filter on P with base /(C/J, and we can form the upper and lower limits of /(H). They are denoted by f(a) = lim/(x) = inf sup/(x) = lim sup/(x), f(a) = lim/(x) = sup inf f(x) = lim inf f(x). x^a [/eu xeU UeU xeU f(x) and f{x) are called the upper and lower limits of f(x) respectively.
40 § 6. Real functions on topological spaces The two concepts which have just been introduced are generalizations of the two well known concepts lim f(x) and lim/(x) of analysis. xn->a x~>a Since the filter corresponding to a net xa->a is finer than the filter of neighbourhoods of a, lim f(x) ^ f(x). For a suitable choice of xa, x<x->a however, equality holds: let /„ be a sequence of neighbourhoods of J (a) with intersection f(a). For each neighbourhood U of a we form the set Un of all xeU for which f(x)el„. The collection of all these Un defines a filter g whose image /'(g) has /(a) as limit. If xa->a is a corresponding net, then lim f{x)=f(a)9 by § 2, 5.(2). (5) The inequalities f(a) ^ f(a) ^ /(a) a/ways /zo/d. (6) If f(x)^g(x) in some neighbourhood of a, then f(a)^g(a) and f(a)^g(a). At every point a for which f(x) + g(x) is defined, we have, by taking limits in (1 a) and (1 b) respectively, (V) (f+g) (a)^ f(a) + g(a), (f+g) (a)^ f\a) + g(a)9 provided that the right hand sides are meaningful. 2. Semi-continuous functions. A real function f(x) on a topological space R is said to be lower semi-continuous at a if f(a)=f(a). It follows from 1.(5) that it is enough to require that f(a)^f(a). A function which is lower semi-continuous at each point is called lower semi-continuous on R. Upper semi-continuity is defined similarly. A function which is both upper and lower semi-continuous at a is continuous at a. Since f\a) is, by the preceding number, the greatest lower bound of the lim f(x\ we have Xot->a (1) f(x) is lower semi-continuous at a if and only if f(a) ^ lim f(x)9 whenever xa->a. *a"a A further criterion is (2) f(x) is lower semi-continuous at a if and only if whenever y <f(a) there exists a neighbourhood U(a) for which y<f(x) for all xeU(a). If the condition is satisfied, we have that y^'mf f(x) for some xeU ' suitable U, so that /(a)^sup inf f(x)=f(a). The converse is also immediately obvious. (3) f(x) is lower semi-continuous on R if and only if for each yeP the set [/> y~\ of those x with f(x) > y is open, or equivalently, if and only if for each yeP the set [/^y] of those x with f(x)^y is closed.
3. The least upper bound of a collection of functions 41 For (2) implies that if a satisfies the inequality f(x)>y, then so does a whole neighbourhood of a. The second assertion follows by taking complements. (4) The sum of finitely many lower semi-continuous functions is lower semi-continuous wherever it is defined. If f(a) + g(a) is defined, then by 1.(7) (f+9) (a)^ f(a) + g(a) =f(a) + g(a). (5) The lower limit f(x) of an arbitrary real function on R is always lower semi-continuous, i.e. f(x)=f(x). For each 8<f(a) there is, from the definition of /(a), an open neighbourhood U0(a) for which 8<f(x) for all xeU0(a). If y lies in U0(a), so does a whole neighbourhood V0(y), so that /(j/) = lim inf f(x)^3 holds throughout U0, and f (a) = \im inf f(y)^S. v xev(y) = c; yecz- Since this holds for every 8<f(a)9 it follows that f(a)^ f(a). It is easy to give examples of semi-continuous functions on any topological space. The characteristic function cp(x) of a subset M of R is defined by cp(x)=\ if xeM, (p(x) = 0 if xeR~M. By (3), M is open or closed in R if and only if cp is lower or upper semi-continuous respectively. For the sets [(p>y] (respectively [(p<y]) can only be the empty set, M (respectively R~M) and R. For any two distinct points x9 y in a Hausdorff space R there is therefore always a finite lower semi-continuous function / with f{x)^f(y). (6) A lower semi-continuous function f(x) on a compact space R takes a minimum value on R. If it only takes finite values, it is bounded below. Since the sets [/^y] are closed, by (3), the non-empty ones among them form a filter base on R. If a is an adherent point of this filter, a lies in all the non-empty [/ ^7], so that f(a)^f(x) for all x in R. A direct consequence is (7) Every continuous function on a compact space takes a maximum and a minimum value on R. If it is finite-valued, it is bounded above and below. (6) and (7) also hold for sequentially compact spaces. The proof of (6) is simple: there is a convergent sequence x„->a in R for which f(xn)-+m9 where m is the greatest lower bound of the values of the function. By (1), f(a) = m. 3. The least upper bound of a collection of functions. If {fp}9 /ie B, is a collection of real functions on the set M, the function which takes the value supf'Ax) at each point x is called the least upper bound sup/^ of the fp. The greatest lower bound inf/^ is defined similarly.
42 § 6. Real functions on topological spaces (1) The least upper bound of a collection fp of functions which are each lower semi-continuous at a point a of a topological space is itself lower semi-continuous at a. In particular the least upper bound of a collection of continuous functions on R is lower semi-continuous on R. We use criterion 2.(2). If y < sup fp(a), y<fp(a) for some /?, so that y< fn{x)fksup fn{x), throughout some neighbourhood U(a). p The next result also follows simply from 2.(2) (2) The greatest lower bound of finitely many functions which are lower semi-continuous at a is lower semi-continuous at a. From Baire's theorem we obtain (3) // / is a lower semi-continuous finite-valued real function on a locally compact or complete metric space R, then the set of points which have a neighbourhood in which f is bounded above is open, and dense in R. For if Mn is the closed set [/^w], every open set 0 is the union of the sets MnnO, and by §4,6.(3) at least one MnnOnO = MnnO contains an interior point. Thus if 0 is some neighbourhood of a point xeR, there is a point y in it which has a whole neighbourhood on which / is bounded above. From (1) and (3) follows immediately (4) If the least upper bound of the lower semi-continuous functions fp on a locally compact or complete metric space R is finite valued, then the set of points possessing a neighbourhood throughout which the fp are uniformly bounded above is dense in R. 4. Continuous functions on normal spaces. The question of the existence of non-constant continuous real functions on topological spaces is more difficult to answer; fundamental for this is the following lemma of Urysohn: (1) A Hausdorff space R is normal if and only if given two disjoint closed sets A and B there always exists a continuous function on the whole ofR with values in [0,1], which takes the value 0 on the whole of A and the value 1 on the whole of B. If the condition is satisfied, the sets [/<i] and [/>i] are disjoint and open, since / is continuous, and contain A and B respectively, so that (N) is satisfied (cf. §3,7.). Suppose conversely that R is normal. We construct open subsets B(p) of R for all dyadic fractions p = k/2n9 fc = 0,l,...,2n, with O^p^l. We set B(0) equal to the empty set and B(l) = R~B.For w = l, applying § 3, 7. (NT), let B(%) be an open set with B(%) => A and B(%) a B{\). If the open sets B(k/2") => A have already been constructed for all fc=l,...,2n, in such a way that B{k/2n) cz B({k+ l)/2") for each k<2\
4. Continuous functions on normal spaces 43 then by § 3, 7.(N') there is always an open set B((2fc+ l)/2" + 1)=>A with S(fc/2")c:B((2fc+l)/2"+1) and 5((2fc+ l)/2"+1) c B((fc+ l)/2"). It follows that £(p) c B(p') whenever p and p' are dyadic fractions with p<p'. We now set /(*)= sup p. Then f(x)= 1 on £, since, if xeB, x lies in x B(p) no£(p)with O^p^l. On the other hand, since, B(p)=> A for all p>0, f(x) = 0 on A The values of f(x) clearly lie in [0,1]. The continuity of f(x) is shown in the following way: let f(x0) = y. If y = 0, let the interval [p,p'~\ be equal to [0,p'] with p' >0, and if 0<y<l let [p,pfli be any interval with dyadic end points p,p' containing y in its interior. Then B(p) is contained wholly in B(p') and B{p')~B(p) is open and contains x0 as interior point. But p<^f(x)^pf for each x in B(p')~B(p), i.e. f(x) is continuous at x0. If, finally, y = l, then x0 lies in R~B(p) for any p < 1; but/(x) ^ p for xeR~B(p\ so that/(x) is also continuous for these x0. If the function f(x) which has just been constructed is replaced by a + (b — a)f(x)9 a,beP, we obtain (2) // A and B are disjoint closed subsets of a normal space R, then there exists a continuous function on R which is equal to a on the whole of A, equal to b on the whole of B, and whose values lie in [a9b~\. Remark. For metric spaces, a function with the properties required in (1) \x9A\ can be given easily. It is enough to put f(x) = -——r^——^t . A topological space is uniformizable (respectively metrizable) if a uniformity (respectively a metric) can be defined on R, for which the corresponding topology is the given one. A first important consequence of (1) is (3) Every normal space is uniformizable. We consider the set C{R) of all finite-valued real continuous functions on R. By § 5, 3. there is a coarsest uniformity $1 on R, with respect to which all the feC(R) are uniformly continuous. It is defined by the vicinities NftC consisting of those (x,y)eRx R for which \f{x)—f(y)\<s. We have to show that the topology X^ defined by this uniformity coincides with the original topology X. Since each / is continuous, each UN(x0)9 which is the set of all y with \f(x0)—f(y)\<e9 is an open neighbourhood with respect to X. If conversely U is an open neighbourhood of x0 with respect to X9 there is, by (1), a continuous function / with /(x0) = 0 and/(y)=l for all yeR~U. The set [/<1] is then a ^-neighbourhood of x0 which is contained in U9 so that X and 3^ are identical. Since every compact space is normal, by § 3, 7. (2), (1) gives information about the existence of continuous functions on compact spaces. It follows from (3) and § 5, 6.(5) that
44 § 6. Real functions on topological spaces (4) Every compact space is uniformizable in one and only one way. For locally compact spaces we have (5) // A is a compact subset of a locally compact space R, and if U is an open set containing A, then there exists a continuous function on R which takes the value 1 on the whole of A, and takes the value 0 on the whole of R~U. Let R' be the compactification of R, as in § 3, 6.(2). The result follows by applying (1) to A and R' ~ U. 5. The extension of continuous functions on normal spaces. A satisfactory account of the existence of continuous real functions on normal spaces is given by Urysohn's extension theorem. (1) Any bounded continuous function on a closed subset M of a normal space R can be extended continuously to the whole of R, preserving the same bounds. Proof. Suppose that f(x) is continuous on the closed subset M of K, and that \f(x)\^c. The subsets Ml = [f^-c/3'] and M2 = [/^c/3] of M are closed and disjoint, so that by 4.(2) there exists a continuous function g1(x) defined on the whole of R, with gx(x)= —c/3 on Mu gl(x) = c/3 onM2and \gl(x)\^c/3 on R. Then if h1(x) is the continuous function on M defined by h1(x)=f(x) — gl(x), |/z1(x)| ^§c on M. Applying the same procedure to h^x), with bound fc, a continuous function g2(x)is obtained on the whole of R, with |#2MI=i'tc> and a continuous function ^2W = ^W~^2W *s defined on M, with \h2(x)\^(^)2c. In general, we obtain for each n a continuous function gn{x) on R with \gn{x)\ ^i(f)"_1-c, and a continuous function hn(x) = hn_l(x)-gn(x) 00 with |/zn(x)| ^(f)"c. The infinite series £ gn(x) converges uniformly n= 1 on the whole of R, and so defines a continuous function F(x\ with 00 I^MI ^i I (!rc=c. OnM, F(x)=f(x)-hl(x)+ X (hM-K+l(x))=\im(f(x)-hn+l(x)). Since \hn+l(x)\^®"+l-c,F(x)=f(x) on M. The extension theorem only holds for normal spaces: If A and B are disjoint closed subsets of R, and if the continuous function defined on AuB by f(x) = 0 on A, f(x)=\ on B may be extended by the extension theorem to the whole of R, then R is normal, by 4.(1). 6. Completely regular spaces. Looking more closely at the proof of 4.(3), it can be seen that the result is still valid, with the following hypothesis about the Hausdorff space R:
7. Metrizable uniform spaces 45 (V) If x0eR and U is a neighbourhood ofx0, there exists a continuous function f(x) on R with values in [0,1] and with f(x0) = 0 and f(x)=l on R~U. A Hausdorff space which satisfies (V) is called completelyregular, oraTychonoff space. Thus (1) Every completely regular space is uniformizable. If x and U(x0) are given, the set [/^i], defined for the function f(x) whose existence is assured by (V), is a closed neighbourhood W of x contained in U, so that every completely regular space is regular. Every subspace of a completely regular space is completely regular, so that subspaces of normal spaces, and in particular of compact spaces, are completely regular. From §3,6.(2) it follows that every locally compact space is completely regular [this also follows from 4.(5)]. Since such spaces are not always normal, the hypothesis of complete regularity is weaker than that of normality. On the other hand the following theorem holds (Tychonoff) (2) Every completely regular space is homeomorphic to a subspace of a suitable parallelotope. Proof. Let {/a}, aeA, be the set of all continuous functions on R with values in [0,1]. By (V), the set of all [/a< 1] forms a basis of open sets in R. Let A be the mapping of R into ^A which sends each xeR to the element yetyA with ya=fa(x). Since R is Hausdorff, for any two distinct x,x' there is always an fa with fa(x)+fa(x'), so that A is one-one. A is continuous, by § 2, 6. (2), and further is open, since the image of each set [/a<l] is open in A(R\ being the complement of the closed set consisting of those y with ya=\. As a special case of (2) we obtain Urysohn's embedding theorem. (3) Every completely regular space with a countable basis, and so for example every separable metric space, is homeomorphic to a subspace of the Hilbert cube, and is therefore normal. For since, by §3,5.(4), there are countably many functions f for which the sets [/)< 1] form a basis of open sets, a homeomorphism is obtained with a subspace of s$w, which by §4,7.(3) is homeomorphic with the Hilbert cube. Normality follows from § 4, 2.(1). 7. Metrizable uniform spaces. Our aim is the converse of 6.(1). In this number we obtain a partial result: (1) A uniform space R is metrizable (i.e. its uniform structure can be determined by a metric) if and only if it is Hausdorff and the vicinity filter $1 of R has a countable base.
46 § 6. Real functions on topological spaces The conditions are necessary, for every metric space is Hausdorff, and the vicinities Nl/n consisting of those (x,y) with \x,y\<l/n form a base of the vicinity filter. The second part is proved in a more general form: a function f(x,y) which satisfies conditions (Dl), (D3) and (D4) of §4,1., and satisfies a weaker form of (D2), namely that /(x,x) = 0 for each x, is called a gauge; we again use the symbol \x,y\. We assert that (2) The uniformity of a uniform space R whose vicinity filter $1 has a countable base can be defined by a gauge. If R is Hausdorff, the gauge is always a metric, and so (2) implies the second half of (1). Proof of (2). If Nl, i=l,2,..., is the countable base of 9t ,we form a sequence of symmetric vicinities Nl9N2,..- for which Nx <= N[ and k N3k+ i^N'kn(] Nh for fc^ 1. The Nt again form a base of JR. t= i We put f(x,y) = mi(\)k, where k runs through all those indices k for which (x,y)eNk; if (x,y)$Nl9 we put f(x,y)=l. Since the Nk are symmetric, f(x,y)=f(y,x); further f(x,y)^.0 and /(x,x) = 0. If (x,y)eNk, (y,y')eNk and (y',z)eNk, then (x,z)eNk cz Nk_i; i.e. from /(x,j;) ^ (i)fc, /(>,,/) ^ (£)* and /(/,z) ^ (i)fc it follows that /(x,z) ^ (£)*" *. We deduce from this that (3) It follows from f(x,y)^c, f(y,y')^s and f(y\z)^c thatf(x,z) ^2e, /or ei;er j; c> 0. We now define a function |x,y| by |x,y| = inf X /(xk-i,xk), fc = 2 where the greatest lower bound is taken over all sequences x1=x,x2,...,xw = 3; of finitely many points of R which begin with x and end with y. \x,y\ is a gauge, for the relations \x,y\ ^ 0, |x,x| =0 and |x,3;| = |3;,x| result immediately from the corresponding properties of f(x,y), and the triangle inequality follows from the definition of |x,y|, since two sequences linking x to z and z to y combine to give a sequence linking x toy. The sets of those (x,y) with f(x,y)^\/2k, k= 1,2,..., form a base for % and so the sets of those (x,y) with \x,y\ ^ l/2fc again form a base for 5R, provided that the relation (4) $f(x,y)^\x9y\^f(x9y) holds.
8. The complete regularity of uniform spaces 47 The second inequality of (4) is trivial. The first is established in the following way: let a sequence x = x1,x2,...,xw = j; be given, and let us set f(xl9x2) + •*• +f(xn-l9y) = M. We shall then prove by induction onn that /(xj)^2M. This inequality is implied for n^4 by (3). Suppose now that the inequality holds for all m<n, n>4. We divide the sequence x1,x2,...,x„_1,xn by taking out a link xh,xh+l in such a way that the inequalities h n 2 /i + 2 hold for the two remaining sequences. By the induction hypothesis we have that f(x,xh) S M and f(xh+l,y) ^ M, and further f(xh,xh+l) ^ M; it follows from (3) that /(x,j;) ^ 2M, so that ?f(x9y) ^ M. Since this holds for all sequences, it also holds for the greatest lower bound of the values of M, i.e. jf{x,y) ^ |x,y|. In the case of a discrete uniform space the construction of (2) gives the metric |x,);| = l for x=\=y. 8. The complete regularity of uniform spaces. Using 7. (2) we now prove (1) Every Hausdorff uniform space is completely regular. If N is a vicinity of R, there is a sequence Nt of symmetric vicinities with Nx cz N, Nf+l c Nt. These define a uniformity on R, which is in general no longer Hausdorff. By 7(2), there is a gauge \x,y\N on R whose vicinities generate the same vicinity filter as the Nt. The vicinity filter N of R is generated by the sets \x,y\N<£, where N is arbitrary in 91. \x0,y\N is a continuous function of y on R, for from \x0,y\N^ \x0,z\N + |<y,z|Arand \x0,z\N^ Ixq^U + Ij^U it follows immediately that ||x0>.yU -|x0,z|.v|^b,zU; thus if \y,z\N<e9 then \\x0,y\N-\x0,z\N\<i:. If U is an arbitrary neighbourhood of x0eR, there exists an N and an £>0 such that \x0,z\N^e for all zeR~U. Since Ixq^I^ is continuous, the function /(x) = Min( 1, —|x0,x|N) is continuous on R. But /(x0) = 0, and f{z)=\ for zeR~U, i.e. (V) is satisfied.
CHAPTER TWO Vector Spaces over General Fields The first three paragraphs are concerned with the elementary and purely algebraic properties of vector spaces E over a general commutative field. In § 7 the lattice V{E) of linear subspaces of E is studied, and § 8 deals with linear mappings from one vector space into another, and their representation by infinite matrices. The problem of the equivalence of these mappings is completely solved. The algebraic dual space £* of all linear functional on E is the theme of § 9. The lattice K(£*) of algebraically closed subspaces of E* turns out to be dually isomorphic with V(E). The end of § 9 is concerned with the most important elementary properties of tensor products of vector spaces. The attempt to establish a complete symmetry between the properties of E and those of £* leads in § 10 to the study of linear topologies on vector spaces. The theory of these linearly topologized spaces is developed in §§ 10—13, following Dieudonne [4], [6], [10], Lefschetz [1] and Mackey [4]. In § 10 we define the concept of a dual system, and introduce the weak linear topology Zls and the topology Xlk defined by the linearly compact subsets of the dual space, between which the original topology lies. As a first application of this theory, § 11 contains the complete theory of the solution of row- and column-finite systems of equations. A simpler constructive method is given for the countable case, which had previously been developed by Toeplitz [1]. § 12 contains the results of Lefschetz about locally linearly compact spaces, and the theory of equations in the countable case, which was developed by Toeplitz and the author (cf. Kothe and Toeplitz [1]), and which is presented in the form given by Dieudonne [4]. The general theory of linearly topologized spaces is continued in § 13, with the introduction of the concepts of linear boundedness and the strong linear topology, and ends with results about strongly reflexive spaces and spaces of countable degree. § 7. Vector spaces 1. Definition of a vector space. Let K be an arbitrary (commutative) field with elements a, /?,..., ^ n, with zero element 0 and identity element 1. A vector space over K (linear space over K, K-module) is a set E with elements (called points or vectors) a,b,...,x, y,... which has the following properties: (LI) For every two elements x,yeE a sum x + y is defined in E; under this addition, E is an abelian group, i.e. for all x,y,zeE we have
1. Definition of a vector space 49 (a) x + y = y + x, (b) x + (y + z) = (x + y) + z9 (c) There exists oeE with x+o = x for all xeE, (d) There exists for each xeE an x'eE with x + x'=o. (L2) For every £eK and every xeE the product £x = x£ of £ with x is defined as an element of £, and for all x,j/e£,^eK we have (e) x(£ + rj) = x£ + xrj, (f) (x + y)£ = xZ + y^ (g) x(£ri) = (x€)ri9 (h) xl=x. We shall establish some simple consequences. (1) For arbitrary a and b in £, the equation a-\-y = b has a unique solution. By (d) there is an a' with a + a' =o; application of (b), (a) and (c) shows that a + {a' + b) = (a + a') + b=o + b = b+o = b9 so that a' + b is a solution. On the other hand if yx and y2 are two elements with a + yl=a + y2, then by adding a! to both sides one obtains first that a' + (a + yl) = (d -\-a)-\-yl = (a + d) + yi=o + yl=y1+o = yl, and secondly that d + (a + y2) = y2> so that yx =y2. It follows from this, by considering (c), that the additive zero element o in E is uniquely determined, and further, from (d), that x' is uniquely determined, x' is written as — x, and the solution b + ( — a) of a-\-y = b as b — a (the difference of b and a). (2) a-0=o,o-<x=o and a( — a)=—{aa) for all aeE, aeK. By (e), a-1 =a(\ +0) = a-1 +a-0. But the equation a-\ +y = a-\ has the unique solution o, so that a-0=o. Likewise it follows from (f) that o-a = (o+o)-a=o-a+o-a, so that o-a=o. Finally, a'<x + a-( — <x) = a(<x — a) = a-0=o, so that a( — a)= —(aa). (3) // x=|=o and a 4=0 then xa4=o. For if x a were equal to o, then by (2), (g) and (h), (x a) a"i = x(a • a~l) = x • 1 = x would also be equal to o. The most important rules of calculation have now been derived; the rules in which sums of n elements are considered, and the associative laws (b) and (g) with more than three elements, are deduced from them in a familiar way, using complete induction. If K is the field P of real numbers or the field V of complex numbers, then E is called a real or complex vector space respectively. 4 Kothe, Topological Vector Spaces I
50 § 7. Vector spaces 2. Linear subspaces and quotient spaces. A subset H of elements of a vector space £ is a vector space provided that whenever it contains x and y it also contains xa + yfi, for arbitrary a, /? in K. H is then called a linear subspace of E. For simplicity we shall usually write o for the subspace consisting only of the zero element o. By a linear manifold in E we mean a subset of E consisting of all elempnts of the form x0 + y, where y belongs to a linear subspace H. We denote this by x0 + H. We also speak of the manifold x0 + H through x0 parallel to H. xl+H = x0 + H for each xlex0-\-H. If xl9...,xn are finitely many elements of £, every element of the form xi<xi+ ••• + x„a„, a,eK, is called a linear combination of *!,..., x„. If M is a finite or infinite subset of £, the collection of all linear combinations of finitely many elements of M forms a linear subspace, the linear span ofM. This can also be defined as the intersection of all the linear subspaces containing M. The linear span of the set {x!,...,*,,} is written as [xx,..., xj. If H is a linear subspace of E and x0 is an arbitrary element of £, the linear manifold x0 + H is also called the //-coset of x0, and is denoted by x0. If we define the sum of two cosets x0 + j)0 to be the coset x0 + j;0, and the scalar multiple x0a to be x0a, the collection of cosets becomes a vector space over K, the quotient space E/H of E by H. The operations which have just been introduced in E/H are defined using particular representatives from the cosets. But if, instead of x0, another element x0 + z0 is taken from the coset x0 = x0 + H, then z0 + H = H, so that the cosets (x0 + y0 + z0) + H and (x0 + y0) + H contain the same elements. The formation of the sum is therefore independent of the choice of representatives. The same is true for multiplication by elements of K, as is immediately verified. Since, further, both operations in E/H are defined in terms of the corresponding operations on the representatives, rules (L1) and (L2) carry over directly to calculations with cosets, and so E/H is a vector space over K. The zero element in E/H is clearly 6 = H. 3. Bases and complements. A finite collection X!,...,x„ of elements of E is called linearly dependent if there is a linear combination xlal + -•• +x„a„, with not all a,- = 0, which is equal to o. In this case at least one of the xt can be written as a linear combination of the others. The elements X!,...,x„ are called linearly independent if they are not linearly dependent; it then follows from a relation x1a1+-- +x„a„=o that all the af = 0. An infinite collection xa, aeA, is said to be linearly independent if each finite collection of them is linearly independent, in the sense which has just been defined.
3. Bases and complements 51 A set {xa}, aeA, of elements of E is called an algebraic basis of E if the xa are linearly independent and if each element of E can be expressed as a linear combination of finitely many xa. It follows directly from the independence of the xa that this representation is only possible in one way, i.e. each xeE can be written as £ xa£a, where only fi- aeA nitely many £a are different from 0, and the £a are uniquely determined by x. We shall leave out the prefix "algebraic" for the time being, provided that there is no possible confusion with other concepts of basis. (1) Every vector space has a basis. Proof. The subsets of E which consist of linearly independent elements clearly satisfy the hypotheses of Zorn's lemma [§2,2.(2)], so that there is a maximal subset {xa} of linearly independent elements of E. Thus if x + 0 is an arbitrary element of £, the set consisting of x and the xa cannot consist of linearly independent elements. There is therefore a linear combination of x and of finitely many of the xa which vanishes, without all the coefficients vanishing. Because of the independence of the xa, the coefficient of x in the linear combination must be different from zero, and so x can be written as a linear combination of the xa. Since, more generally, the subsets of E which consist of independent elements and which contain a fixed independent set {yp} also satisfy the hypotheses of Zorn's lemma, we have (2) Every system of linearly independent elements of E can be extended to a basis of E. Two linear subspaces G and H of E are called algebraically complementary to each other if each xeE can be represented in one and only one way as a sum x = y + z, with yeG and zeH. If elements yp form a basis for G and elements zy form a basis for H, then the set consisting of the yp and the zy is clearly a basis for E. (3) Every linear subspace G of a vector space E has a complement. Proof. Let {yp} be a basis for G and let {xy} be a basis for E/G. If an element zy is picked out of each of the cosets xy, the yp and zy together form a basis of E: if xe£, then x = £xy£y, so that x = £zy£y + );, yeG. The expression x = YJzy€7 + Y,yprlp then follows from y = Yjyprjp. On the other hand the yp and zy are linearly independent, for from YJypr1p + YJzy£y=0 it follows, by going over to cosets, that ^xy^y=6. Hence £y = 0 for all 7, and so ^ = 0 for all /?. The linear span H of the zy is clearly a complement for G. (3) can also be proved by applying (2) to a basis {yy} of G. We observe that the complement to G is not in general unique, since the zy can be chosen arbitrarily in the xy. Further let it be stressed 4+
52 § 7. Vector spaces that we have only given existence proofs, depending upon the axiom of choice, for (1) and (2). We return to constructive methods in the case of a countable basis in § 11. 4. The dimension of a linear space. A basis {xa}, aeA, of E has a certain magnitude, given by the cardinality of A—the "number" of elements of the basis. We show that this number is the same for different bases of E. n (1) n+\ linear combinations yt= £ xk<xkh i=l,...,n+l, of n k= 1 elements xkeE are always linearly dependent. This is clear for n= 1. Suppose that the result is true for n— 1, and that further an4=0 (which we can assume without loss of generality). <xli Then by the induction hypothesis, the n elements y\ = yi — yx —, an i = 2, ...,n+l, are linearly dependent, being linear combinations of x2,..., x„, and so we can write n+ 1 (2) Z /,/?,=<>. 1 = 2 But then the yt are also dependent, for (2) implies the relation n+1 n+ 1 E M-y. E — P<=° r = 2 i = 2 a 1 1 and not all the /Jf vanish. If now £ has a basis with finitely many elements, let xl9..., xd be a basis with smallest possible d. If {y(i} is a second basis with cardinal / then, by (1), / must be less than or equal to d, since on the one hand the yp are linear combinations of the xh and on the other they are linearly independent. If the bases of £ are all infinite, we finish the proof as follows: Let {xp} and {ya} be two bases with cardinality d and / respectively. There are equations (3a) Xp = Y,ya^afn (3 b) )><r = Z Xp P Po ■> a p in which for each p or a there are only finitely many non-zero <xap or fipa respectively. In the equations (3 a) each ya has at least one coefficient a(Tp4=0. Indeed if all the aaoP were zero for some <70, then we would have >'*„ = E xpPpo0 = E (E x* a° p ) Pp°o = E ^ E a-P /^p^o /; p \ <t / a p
5. Isomorphism, canonical form 53 which expresses yao as a linear combination of finitely many ya>, o' 4=t70, contradicting the linear independence of the ya. Using (3 a), assign to each xp the set Mp of those finitely many ya for which <xap^0. The set M of all the vectors ya is the union of these d sets Mp, so that the number / of ya is less than or equal to X0 d, where X0 is the cardinal of the set of natural numbers. By hypothesis, d is an infinite cardinal; by a well-known theorem of set theory it follows that tf0d = d. f=d then follows from the inequality f^d and the corresponding inequality d^f. We have therefore shown (4) Two different bases of a vector space E always have the same number of basic elements. This cardinal is called the algebraic dimension d(E) of E. 5. Isomorphism, canonical form. Two vector spaces Ex and E2 over the same field K are called algebraically isomorphic if there is a one-one correspondence x<-»x' between the elements of Ex and the elements of £2> under which (x<x + yP)' = x'a + y'P, for all x,yeEi and all a, /?eK. We shall use the symbol Ex = E2 for this. The result of the preceding number can now be expressed as follows: (1) Two vector spaces Ei and E2 over the same field K are isomorphic if and only if they have the same dimension. Since an isomorphism sends a basis of Et into a basis of E2, the dimensions must coincide. On the other hand if they are equal, then there is a one-one correspondence xa<-»x'a between a basis of £x and a basis of E2, which defines an isomorphism x = £xa£a<-»£x/a£a = x/. To each cardinal number d there is, up to isomorphism, at most one vector space over K of dimension d. On the other hand a vector space of dimension d over K is obtained in the following way: Let a run through a set A of magnitude d. If a £aeK is made to correspond to each aeA, and if only finitely many £a are non-zero, then we call the function defined in this way a finite vector x={^a}, with d coordinates £a. As usual we define s + n as the vector with coordinates ^a + na, and further define xp, peK, as the vector with coordinates £ap; the set E of these vectors is then a vector space over K. The null-vector o with all coordinates vanishing is the zero element of E. We denote by ea the a-th unit vector, whose a-coordinate is 1, and whose other coordinates vanish. Clearly ac = ^ea^a, and the ea are a linearly independent and therefore form a basis for E. We call the space E obtained in this way a d-dimensional finite coordinate space cpd(K) over K. If d = X0, we write simply cp(K).
54 § 7. Vector spaces In this expression the index set A does not appear explicitly, and it can always be replaced by one with the same cardinality. If a uniquely determined canonical form is required, it is natural to take the set Qd of those ordinals with smaller cardinality than d. This corresponds to the practice, in the finite and countable cases, of taking natural numbers as indices, and has the advantage that one can fall back upon the order of Qd in, problems for which an ordering of the coordinates is useful. Let it be stressed, however, that this order is not essentially connected with the concept of basis, but is introduced in addition. By (1), an isomorphism is obtained between a general d-dimensional vector space E and cpd, when the elements xa of a basis of E are made to correspond to the unit vectors ea of cpd. 6. Sums and intersections of subspaces. A linear subspace of E is always obtained when the set theoretic intersection f] Fa of a collection a of linear subspaces Fa of E is formed. The intersection of a collection of linear manifolds is either empty or a linear manifold. If in a corresponding way the union \J Fa is taken, it need not in general be a linear SUb- space, although its linear span is. This linear span is called the sum £ Fa a of the Fa. For general subsets Ma of F, £ Ma denotes the set of all finite sums £ xa.(xa.eMJ. For linear subspaces Ma, this is again the i=\ linear span of the Ma. The plus sign is used as well, when there are finitely many summands, e.g. Ml+M2. The sum F = £ Fa is said to be direct, if every xeF can be written in only one way as £ xa, xaeFa (of course, there are only finitely many xa=|=o). For direct sums we use the symbols © Fa and Fi®F2 re- a spectively. An example of a direct sum is given by two complementary subspaces G and H of E; E = G®H. (1) The sum £Fa is direct if and only if Fan £ Fp=o, for each a. If the condition is satisfied, it follows from £xa=o that x(X= £ ( — xp)9 and so x^eF^n £ Fp; this implies that xa=o for all a, so that the representation x = £ xa as a sum is unique. a On the other hand, if this is the case, and if zeFan £ Fp, then the representation o = z — z, with zeFa and —ze £ Fp, must coincide with o=o+o, so that z=o. /J + a
7. Dimension and co-dimension of subspaces 55 (2) If F and G are two linear subspaces of E, and if Fx and Gx are complementary to F r\G in F and G respectively, so that (3) F = Fl®(FnG)9 G = Gl®{FnG), then (4) F + G = F1®Gl®{FnG). In this way the sum F + G is reduced to a direct sum. Proof. Clearly F + G = F1 + G1+(FnG). But this sum is direct, for it follows from x + y + z=o, xeFl9 yeGl9 zeFnG, that x = —y — zeG and so that xeF1nG = F1n(FnG)=o, which shows that x=o. But then y = z=o must hold, by (3). It follows from (2) that (5) If F and G are two linear subspaces of £, there is a complement H of Ffor which G = (GnF)®(GnH) holds. For if G: is defined by (3) and if L is a complement of F + G in £, H = G1®L has the required properties. The next two isomorphism theorems, well-known in group theory, follow from (2). (6) If F and G are two linear subspaces of £, then (F + G)/G ^ F/(F n G). (7) If F and G are two linear subspaces of £, with F cz G, then E/G^(E/F)/(G/F). Proof. By (3), F/(FnG)^Fl9 and by (4), (F + G)/G^Fl9 which establishes (6). If further F cz G cz £, then we can write E = F®F1®H9 where F1 is a complement of F in G and if is a complement of G in E. E/G^H and E/F = Fl®H; under this isomorphism G/F is mapped onto Fl9 so that (E/F)/(G/F) is also isomorphic to H. 7. Dimension and co-dimension of subspaces. Let G be a linear sub- space of £, and let if be a complement of G in E. If each ze H is made to correspond to its coset z in the quotient space E/G, an isomorphism is established between H and E/G9 so that (1) All the complementary subspaces H of a linear subspace G of E have the same dimension, namely the dimension of E/G. The dimension of E/G is called the co-dimension, or the defect, c{G) of G in E. We have (2) d(G) + c(G) = d(£), for a basis of G and a basis of one of its complementary spaces together form a basis of E.
56 § 7. Vector spaces More generally, interpreting the sums as sums of cardinal numbers, we have (3) If E=® Fa, then d(E) = £d(Fa). a a (4) // F and G are two linear subspaces of F, then (5) d(F + G) + d(F nG) = d(F) + d(G), (6) c(F + G) + c(FnG) = c(F) + c(G). By going over to dimensions, equation (5) can be read off from equations (3) and (4) of the preceding number. If, further, H is a complementary subspace of F + G in F, then we have by 6. (4) y y) E = H®F1®G1®FnG. (6) follows by comparing the dimensions of the complements of FnG, F, G and F + G which appear in this decomposition. If a linear subspace H has codimension 1 in F, both H and manifolds x0 + H parallel'to H are called hyperplanes in F. (7) Every linear manifold is the intersection of the hyperplanes containing it. It if sufficient to show this for a linear subspace F. Suppose that z does not lie in F. The linear span [z] of z consists of all multiples z £, qe K. The sum G = F + [z] is direct; thus if H is a complement of G, E = F®\_z~\®H. Clearly F®H has codimension 1 in F, and it is therefore a hyperplane, which does not contain z. Since z is an arbitrary element not lying in F, the intersection of all the hyperplanes containing F is equal to F. 8. Products and direct sums of vector spaces. Given vector spaces Fa, aeA, the following is a simple method of constructing a new one. We form the set-theoretic product F = TT Ea (cf. § 1, 8.). F becomes a a vector space when we define x + y to be the function x^ + y^ on A, and define x£, £eK, by (x£)a = xa£. F is called the product oftheFa. If all the Fa are equal to F, we also write FA for the product, or even Fd, where d is the cardinality of A, when it does not matter going over to an isomorphic space. In particular if F is equal to K, Kd is the vector space of all vectors x={£a}, with d arbitrary coordinates in K: this is also called the linear coordinate space cod(K); in the case where d = K0, we simply write co(K). As with (pd{K\ we may take the ordinals of Qd as coordinate indices, if necessary. Let us remark straight away that <jod need not have dimension d (cf.§9,5.).
9. Lattices 57 The collection F of those xeE = TT Ea with only finitely many non- a zero xa again forms a vector space. This is clearly equal to the direct sum © £a, where Ea is the vector space, isomorphic to £a, which a consists of those xeE with xp=o for all /}=}=a. We write © Ea for F, ^ a identifying £a with £a, and again call © £a the direct sum of the £a. a If all the £a, aeA, are equal to K, then © Ea is just the space cpd(K) a which has already been introduced in 5., where d is the cardinality of A. © £a is a subspace of TT £a; they coincide if and only if the set A a a of indices a is finite. In this case either expression Ex x ••• x En or E1 © • • • © En can be used. A further method of forming products, the tensor product of vector spaces, is dealt with in § 9, 6. 9. Lattices. In 6., we introduced the two operations + and n on the linear subspaces of E. We can understand their properties most clearly by using the concept of a lattice. For this we refer to the ideas of §2. A partially ordered space Kis called a lattice if every set consisting of two points a, beV has a least upper bound c and a greatest lower bound d in V. c is called the union avb of a and b, d the intersection a/\b of a and b. A lattice V is said to be complete if it satisfies the stronger condition that any arbitrary set {aa} of elements of V has a least upper bound and a greatest lower bound. These elements are again called the union \y aa and the intersection /\ aa respectively. a a It follows easily from the definition of the operations v and a that they are commutative and associative (even for infinitely many terms) and further that they are monotonic with respect to the partial order, i.e. (1) // a^b, then avc^bvc and aac^ac. As an example, let us mention the topologies on a set, which, by §1,6., form a complete lattice with respect to the relation "finer". Two lattices Vx and V2 are said to be isomorphic if there is a one-one correspondence a1*-+a2 between the elements of V1 and the elements of K2, under which al^bl if and only if a2t^b2. Unions and intersections of corresponding elements also correspond. Two lattices Vx and V2 are said to be dually isomorphic if there is a one-one correspondence ax <->a2 under which ax^bx if and only if b2^a2. The union of a collection of elements then corresponds to the intersection of the corresponding elements, and conversely.
58 § 7. Vector spaces To every lattice there corresponds its dual, obtained by interchanging rg and ^. If there are elements 0 and 1 in a lattice V with O^ga and a^ 1, for all aeV, then they are called the zero element and unit element of V respectively. Every complete lattice has a zero and a unit element. A lattice is said to be modular if it satisfies the condition (2) // a^c, then a v(b Ac) = (a v b) ac, and is said to be distributive if it satisfies the stronger condition (3) a v (b a c) = (a v b) a (a v c), a a (b v c) = (a a b) v (a a c) for all a, b and c. A lattice V with a zero and a unit element is said to be complemented if (4) For each a there is at least one a' with ava' = l,aAa' = 0. A complemented distributive lattice is called a Boolean algebra. If the symbols v and a are interchanged in (2), (3) and (4) and if ^ is replaced by ^ in (2), then the statements remain unaltered, and so these properties are preserved under isomorphisms and dual isomorphisms. 10. The lattice of linear subspaces. We show (1) The linear subspaces of a vector space E form a complete complemented modular lattice V(E) under the relation A<^B. The lattice operations \J and /\ are just £ and (°), respectively. a a a a Proof. It is immediately clear that V(E) is a complete lattice. To say that it is complemented is to say that each A has a complementary subspace. It remains to show that V(E) is modular. Suppose that A <= C. If x is an element of (A + B) n C, then, considered as an element of A + B, it can be written as x = y + z, ye A, zeB. Since xeC and yeA^C, it follows that zeC, and so zeBnC. Thus xeA + (Bc\C). We have therefore shown that (A + B)nC<=A + (BnC). On the other hand the following holds in any lattice, as is easy to see: (2) // a ^ c, then a v (b a c) ^ (a v b) a c. From this the proposition follows. V(E) is not distributive, for although, as in any lattice, one half of the distributive law holds, namely (3) C\{F.+Gf)z>(C\Fa)+(C\Gf), £(^G,) = (XF.)nteG,), <x,P a P a, P a P
1. Definition and rules of calculation 59 equality does not always hold in (3), even for the linear subspaces of a finite-dimensional E (example!). We also remark that, using 6.(1), the idea of direct sum can be defined using + and n, and so can be defined in terms of the lattice. § 8. Linear mappings and matrices 1. Definition and rules of calculation. Let E and F be two vector spaces over K. A correspondence A, which sends each xeE to an AxeF, is called a linear mapping, a linear transformation or a homomorphism of E into F if (1) A(xz + yP) = (Aa)aL + (Ay)P holds, for all a, /Je K and all x, yeE. The mapping which sends each xeE to the zero element of F is denoted by 0. The sum A + B and the product A a of mappings of E into F are defined by (2) (A + B)x = Ax + Bx, {Aot)x = (Ax)(x. Clearly we have (3) The set &(E,F) of linear mappings of E into F forms a vector space over K. If B maps E into F and A maps F into G, then the formula (AB)x = A(Bx) defines a linear mapping of E into G, the product AB. This product is associative, the two distributive laws A(B + C) = AB + AC, (A + B)C = AC + BC both hold, and by (1) we have (AB)ol = A(B(x) = (Ax)B. A linear mapping of E into itself is also called an endomorphism. We write 8(E) for the set of all endomorphisms of E. The identity endomorphism is denoted by /. A vector space jR over K is called an algebra over K if a product ab is defined for any two elements a and b of jR, which satisfies the rules (4) (ab)c = a(bc), (5) a(b + c) = ab + ac, (a + b)c = ac + bc, (6) (a b) a = a(b a) = (a a) b. With this terminology we have (7) The set 8(£) of endomorphisms of a vector space E over K is an algebra over K, with I as multiplicative identity element.
60 § 8. Linear mappings and matrices 2. The four characteristic spaces of a linear mapping. Let i be a linear mapping of E into F. The set A(H) of images of a linear subspace H of £ forms a linear subspace of F. In particular, A(E) is called the image space of A The set of all yeE with Ay=o forms a linear subspace of F, the null-space or kernel N\_A~\ of the mapping. If N\_A~]=o, then A is one-one, and A is called a monomorphism of E into F. In agreement with § 7, 5., A is called an isomorphism of E onto F if N\_A~\=o and 4(F) = F. An isomorphism of E onto itself is also called an automorphism. The automorphisms of F form a group, the linear group A(F) of F. If H is a linear subspace of F, the monomorphism Je©(//,F) which sends each yeH to the same element j;, now considered as an element of F, is called the embedding (injection) of H in F. If A(E) = F, A is said to be an epimorphism of F onto F. If H is a linear subspace of F, then the mapping K which sends each xeE to the coset x in E/H is an epimorphism of F onto F///, which we call the canonical mapping of F onto E/H. With this terminology we have (1) Every linear mapping A of E into F is the product JAK of the canonical mapping K of E onto E/N\_A~\, an isomorphism A of E/N\_A~\ onto A(E), and the embedding J of A(E) into F. The mapping A = J A is a monomorphism of E/N \_A] into F. We call E/N\_A~\ the inverse-image space of A. By §7,7.(1) N\_A~\ has complementary subspaces C/[>4], which, it is true, are not uniquely determined, but which are all isomorphic to E/N\_A~\. A then maps each such space U\_A] isomorphically onto A\_E~\. U\_A~\ will also be called an inverse-image space of A. The fourth characteristic space for A is F/A{E). The complementary subspaces of A(E) in F are again isomorphic representatives of F/A(E); these we call complements C[4] of the image space, as indeed we call F/A{E). 3. Projections. If E = E1® E2 is a decomposition of F into complementary subspaces, then each x in F can be expressed uniquely in the form x = x1+x2, with x1eE1 and x2eE2. If we send each x to its component xx in F1? then this mapping is linear, it maps Ex identically onto itself and it sends F2 to zero. This endomorphism is called the projection PEl of F onto Fl5 with null-space F2, or in the direction F2. P£2j = PEl, so that the endomorphism is an idempotent element of the algebra ©(F). If conversely P is an idempotent endomorphism of F, so that P2 = P, then it follows from P(Px) = P2x = Px that the subspace E1=P(E) is
4. Inverse mappings 61 mapped identically onto itself by P. The set of all x — Px forms a linear subspace E2 of F, which is sent to zero by P. The decomposition x = Px + (x — Px) = xi+x2 is unique, for from x1+x2=o, x1eEl, x2eE2, it follows, by multiplying by P, that xl=o. E2 is therefore complementary to F:. As can be seen immediately, / — P is the projection of £ onto F2, with null-space El9 and P(I — P) = (I — P) P = 0. Thus we have established (1) To every complementary decomposition E = EX® E2 there correspond two projections PEl and PEl, with El=PEl(E\ E2 = PEl{E\ PElPEl = PE2PEl = 0 and PEl + PE2 = L Conversely every idempotent mapping Pe(B(E) determines a complementary decomposition E = P(E)®(I — P)(E). As a first application we show (2) If E= ® Fa, then <S(E,F) is isomorphic, as a vector space over K, to TT S(Fa,F). " a Let Pa be the projection of E onto Fa with null-space © Ea>. Then a' =t= a for each xeE we have the representation x = £xa = £Pax. If /4e8(E,F), then we have « (3) Ax = YJA{Pax) = YJ{APx){Pax). <x <x 4Pa, considered as a mapping from Fa in F, is denoted by Aa. Thus, by (3), to each A there corresponds a vector {/la}eTT ®(Ea,F). Conversely, given such a vector {yla}, the formula (4) Ax = YjAa{Pax) <x defines a linear mapping in 8(E,F). 4. Inverse mappings. If /1g6(F,F) and 5e®(F,F), and if BA = IE, where /£ is the identity mapping of F, then £ is called a left inverse or left reciprocal of A. Similarly Ce<B(F,E) is called a right inverse of A if >4C = /F, where IF is the identity mapping of F. If Be&(F,E) is a left inverse of >4, and if Ce<B(F,E) is a right inverse of A, then B = BIF = B(AC) = (BA)C = IEC = C. In this case we speak of the (two-sided) inverse of A, and denote it by A'1. A mapping A which has an inverse A~J is said to be invertible. (X) A mapping /4gS(F,F) is invertible if and only if it is an isomorphism of E onto F. For if A is an isomorphism, the inverse A{~1] is a mapping of F onto F with Ai~1)A = IE and AA{~l) = IF. On the other hand if A is
62 § 8. Linear mappings and matrices invertible, A is one-one, for it follows from Ax=o that A~l(Ax) = (A~lA)x = x=o. Further, A(E) = F, for given yeF, y = (AA~l)y = A(A~l y), so that y is the image of A'1 yeE. If A is not an isomorphism of E onto F, then either A(E) is a proper subspace of F, or A is not one-one, or both, so that the inverse A{~1] is either defined only on a proper subspace of F, or is not a point-mapping of F into E. In neither case, therefore, is it a mapping in <5(F,E). We now investigate the extent to which it is generally possible to invert A, using a suitable mapping in 8(F,F). By 2., we can choose an inverse image space If [4], complementary to the kernel N[>4], and a space C\_A~\ complementary to A(E). A maps U\_A] isomor- phically onto A(E). This correspondence can be inverted in a unique way. If we further stipulate that the elements of C\_A] are mapped into the zero element of F, then a linear mapping B is defined on the whole of F, with kernel C\_A], inverse image space A(E) and image space £/[/4]. If we combine A and B, BA turns out to be the projection of F onto £/[/4], with null-space N[/4], whereas A B is the projection of F onto A(E\ with null-space C\_A]. Conversely, by 3.(1), every projection PN[A] determines a specific £/[,4], and every projection PA(E) determines a specific C\_A~\, and so we arrive at the following conclusion: (2) // Ae<5(E,F), if PN[A] is a projection of E onto the null-space of A and if PA{E) is a projection of F onto the image space of A, then there is a mapping Be&(F,E) with (3) BA = IE-PN[A], AB = PA{E). Using this, we can answer the question of when A has a left or right inverse. (4) Let A^O be a linear mapping of E into F. Then one of the four following cases holds: 1. N\_A~\=o, A(E) = F. A is invertible. 2. N\_A] #=o, /4(F)#= F. A has neither a left nor a right inverse. 3. Af[y4]4=o, A(E) = F. A has no left inverse, and has at least two right inverses. 4. N\_A~\=o, A(E)^F. A has at least two left inverses, but has no right inverse. Proof. A can only have a left inverse if N(A)=o, and can only have a right inverse if A(E) = F. The conditions are therefore necessary. But if they are satisfied, then the existence of left and right inverses follows from (3). The fact that there is more than one inverse in cases 3.
5. Representation by matrices 63 and 4. follows from the fact that there is more than one projection PN[A] and PA(E) respectively. Examples. If E is n-dimensional, then cases 3. und 4. cannot arise for endo- morphisms of £, as is well-known. If, however, we consider the identity mapping of <p„(K) into itself as a mapping of (pn(K) into (/>„,(K), with m>n, we obtain an example of the fourth case. If K has only finitely many elements, then in this example there are only finitely many left inverses. If K has infinitely many elements, then there are always infinitely many inverses in cases 3. and 4. If x0,*!,... is a basis of cp{K\ then the mapping x,—>xf+1, i = 0,l,..., is an example of case 4., and the mapping x0->o, xI+1->xf, i = 0,1,..., an example of case 3. 5. Representation by matrices. It is easy to obtain a general picture of all the linear mappings of E into F. To this end, we choose bases {xv}, veN, and {yM}, jueM, in E and F respectively. If we specify an arbitrarily chosen zveF as image, for each xv, then the correspondence x = Yaxv£>y-*y = Zzv£v defines a linear mapping A, and conversely V V every linear mapping is determined by the images Axv = zv, in this way. Each zv can be expressed in terms of the basis y^ of F (1) zv = E)VaMv> °eMVeK, veN. If we combine the elements Axv, veN, to form a vector {Axv}, then A is given by the equation (2) M*v} = {X3^aMvj- The mapping is therefore fully determined by the coefficients oeMV. Following normal practice, we call the collection of terms aMV a matrix 9l = ((oeMV)), jugM, vgN, with elements in K, and defined over M x N. The terms oeMoV, veN, form the ju0-th row and the terms oeMVo, /*eM, the v0-th column; the elements of M and N are called row- and column-indices respectively. It follows from (1) that every column of our matrix 21 has only finitely many non-zero elements, and 91 is therefore said to be column- finite. If d and e are the dimensions of E and F respectively, then 91 has e rows and d columns. If 9I = ((aMV)), ^gM, vgN, is a matrix, and if x = {^}, /igM, is a vector over the set of row indices of 91, then the product t) = s9I is defined as the vector with components ^v = Z^MaMv provided that each of these sums has only finitely many non-zero summands. In a corresponding way 913 is defined as the vector with components XaMvCv> when 3 = {Cv} is a vector defined over N. V
64 § 8. Linear mappings and matrices With this terminology, (2) becomes (3) {Axy} = {y,}W. Given x = £xv£ve£, it follows from (2) that (4) Ax = £04xvKv = X (Ejw) ^v = X3V ZaMv^v, v v n n v or, in terms of matrices and vectors, (5) Ax={Axv}x = ({yJVL)x={yJ(VLx). If x is replaced by the vector x with respect to the basis xv in £, then A x is represented by 21 s with respect to the basis y^ in F. Thus we have shown (6) // bases {xv}, veN, and {y^}, /ieM, are chosen in E and F respectively, then using (3), every linear mapping y = Ax from E into F corresponds to a column-finite matrix 2I = ((aMV)), /zgM, vgN; if x is the vector corresponding to x with respect to the basis {xv}, and n the vector corresponding to y with respect to the basis {y^}, then the expression n = 21 x is the representation of A. Conversely any column-finite matrix with e rows and d columns can be interpreted as representing a linear mapping from a d-dimensional vector space into an e-dimensional vector space. Let B be a linear mapping of F into G, and suppose that, for a basis {z;}, AeA, we have (7) By, = Y,zxPx,- X Then for the compound mapping BA we have (8) (BA)xv = B(Axv) = £(Bjga„v = X (Y zJXf\ a„v = £ zx £ /^a„v. H H X ' A. \i The matrix of BA is therefore the matrix ((X/Ja^v))' AeA, veN, which is written as the matrix product 9321; the elements of it are the scalar products of rows of 93 with columns of 21. The associative law for mappings goes over directly to the representative column-finite matrices, so that we always have (£(93 21) = ((£ 93)21 The sum A + B of two mappings in <B(E,F) corresponds to the sum 21 + 93 = ((aMV + j8MV)) of the corresponding matrices with respect to the same bases, and the mapping Ap, peK, corresponds to the matrix 2Ip = ((oeMVp)). The mapping 0 is represented by the zero-matrix D, all of whose elements are zero, and the identity mapping / by the unit matrix g = ((eMV)), with eMM=l and eMV = 0 for fi + v.
6. Rings of matrices 65 6. Rings of matrices. A matrix 9I = ((aAiV)), fief\A, veN, is said to be square if M = N. A matrix with the same number of rows and columns need not be square, but it can always be changed into a square one by modifying the indices. The operations 91 p, $1 + 93 and 93 $1, introduced in 5. for column-finite matrices, may also be defined for general matrices, although it should be observed that the product 93 91 can only be defined if the set of column-indices of 93 coincides with the set of row-indices of 91, and if each sum X/^aMv nas onh finitely many non-zero summands. ^ A set 9JI of square matrices over the index set N, with elements in K, is called a linear matrix ring over K if SR forms an algebra over K with respect to the matrix operations 9lp, 21 + 93 and 9321. Two algebras R and R' over K are said to be isomorphic if there is a one-one correspondence a<^>d between the elements of R and R' for which (1) (a + b)' = a' + b\ (ab)' = a'b\ {ap)' = a'p holds, for all a, beR and all peK. They are said to be anti-isomor- phic if the one-one correspondence satisfies the equation (ab)' = b'a' instead of the second equation in (1). We consider the endomorphisms A of a d-dimensional vector space E. If {xv} is a basis of E, we can take the set Qd of ordinals of magnitude smaller than d as index set. We take the same basis in F = E, so that the matrix corresponding to A, as in 5.(3), is square. The results of 5. then become (2) The algebra ®(E) of endomorphisms of a d-dimensional space E over K is isomorphic to the linear matrix ring 2RJ(K) of all dxd square column-finite matrices over Qd, with elements in K. A linear matrix ring SR over K, whose square matrices 91 are defined over the index set N, is said to be maximal, if there is no linear matrix ring SRj over K, whose matrices are again defined over N, which contains SR as a proper subset. (3) 2RJ(K) is maximal. It can clearly be supposed that d^K0. It is not possible to extend SRJ(K) by adding a matrix (£ which is not column-finite: if, for instance, the v-th column of (£ contains infinitely many non-zero elements, than the dxd matrix 21 which has the v-th column of & as v-th row, and otherwise has zero rows, belongs to 2RJ(K). The scalar product of the v-th row of 21 with the v-th column of & cannot be formed, since the sum must be taken over infinitely many non-zero terms. 5 Kothe, Topological Vector Spaces I
66 § 8. Linear mappings and matrices 7. Change of basis. If {x'v}, v'eN', is a second basis of E, then it is related to the basis {xv}, veN, by equations (la) {xv} = {<,}£, <5, = ({yv.v)), VeN', veN, (lb) {x'v,} = {xv}X), 3 = ((<5VV0), veN, v'eN'. (£ and D are column-finite, and by putting (la) in (lb) and conversely we obtain £D = (£N,, D(£ = (SN, the unit matrices over the sets N' and N respectively. A column-finite matrix £ = ((£vv/)), veN, v'eN', is again called a left inverse of the column-finite matrix 2I = ((av,v)), if £2I = (£N, and similarly a right inverse ?) = ((f/VV')) is defined by 21*2) = (£N,. As in 4., it follows from the existence of both a right and a left inverse of $1 that they coincide, and are uniquely determined. In this case 91 is again said to be invertible, and the two-sided inverse is denoted by 9I"1. Thus in the present case D = (£_1. Conversely, given a column- finite invertible matrix (£ with two-sided inverse D, it is easy to see that a change of basis of E is given, using (1 a) and (1 b). We remark that £ is square only if N = N'. If x is an element of E, we obtain from (1 a) that (2) x={xv}x = {x'v,}£x, so that (3) Changing the basis {xv} to the basis {x'v>} = {xv}&~1 changes the representative xto&x. If we also change the basis {y^}, /ieM, of F to a new basis {y^} = {<yAi}95~1, /i'eM', then, using (lb) and 5.(3), we get the following for the matrix corresponding to A: (4) {xx;.} = Mxv}c-1 = {^}9rc-1={y,.}®sie;-1, so that (5) // the bases {xv} and {y^} are changed to the bases {x'v>} = {xv}(£-1 and {y,fl'} = {yfl}^B~1 of E and F respectively, then A is represented by ESICrJ instead of by VL. In the case where F = E, if we take the same change of basis both for the original elements and for the image elements, then from (5) we get (6) Changing the basis {xv} to the basis {xv,} = {xv}&~1 changes the representative 21 of an endomorphism Ae&(E) to the representative CSIGT1. 8. Canonical representation of a linear mapping. We continue with the ideas of 2. We call the dimension of the image space A(E) the rank r(A) of A, the dimension of the kernel N[A~\ the nullity s{A) and finally
9. Equivalence of mappings and matrices 67 the dimension of F/A(E)9 and so of a complement C[A] of the image space, the defect, or co-nullity, s'(A). The rank r(A) is clearly also equal to the dimension E/N\_A~\. In order to represent A by a particularly simple matrix, we proceed in the following way. We choose an inverse image space £/[A], and then choose a basis zv>, v'eN', of this space. By adding a basis zv», v"eN", of N\_A~\ to this, we obtain a complete basis of E. The elements wv, = Azv> then form a basis of A(E), and by adding a basis wM», /i"eM", of some C[A] to this we obtain a complete basis of F. A is then given by (1) /lzv, = Hv(v'eN'), 4z,.=o(v"6N"). For the given bases of E and F, X is therefore represented by a matrix £), for which the elements <5vV, v'eN', are equal to one, and all the others are zero. Thus we have shown (2) By choosing bases suitably in E and F, the linear mapping A of E into F is represented by a matrix £), which reduces to an r(A) x r(A) square unit matrix when its s'(A) zero rows and s(A) zero columns are deleted. T) is called a canonical representation of A, in the wide sense. If it is required to use given sets N and M as index sets of the bases in E and F respectively, then N' and N" can be chosen as complementary subsets of N, but for the elements Azv> it is only possible to use indices /i'(v')eM' cz M which are in one-one correspondence with the indices v'. The canonical representation T) in the narrow sense which is obtained in this way has <5 ,(v,)>v, = 1, and so the deletion of the zero rows and zero columns leads to an r(A) x r(A) matrix, which has exactly one one in each row and column. In order to obtain a canonical representation of an endomorphism ,4 e 6(F), it is necessary not just to take one basis of F, but to take two different bases, one for the elements x, and one for the elements A x. 9. Equivalence of mappings and matrices. Two linear mappings A1 and A2 in 6(F,F) are said to be equivalent if there exist an invertible mapping B in 6(F) and an invertible mapping C in 6(F) for which A2 = CA1B. The equivalence defined in this way is reflexive, symmetric and transitive, so that the mappings in 6(F,F) fall into classes of equivalent mappings. If we choose bases xv and yM in F and F, then for the corresponding row-finite matrices we also have $I2 = ^^i®> where 33 and (£ are invertible and square. In this case we say that $It and $I2 are equivalent in the narrow sense, and the equivalence of the mappings implies the equivalence of the matrices in the narrow sense, and conversely. 5*
68 § 8. Linear mappings and matrices For this matrix equivalence it is necessary for 2^ and 2I2 to t>e defined with the same row and column index sets, respectively. More generally, if $It and $I2 are two column-finite matrices with the same number of rows and the same number of columns, an equivalence in the wide sense can be defined by 9l2 = ^^i®> where now the requirement is that 33 and (£ are invertible. If 2^ and 212 have the same index sets, this is the same concept of equivalence as before, as 33 and (£ must then be square. Theorem (5) of 7. can now be expressed in the following terms: (1) The representative matrix of Ae&(E,F) obtained by changing the bases in E and F is equivalent to the original one, in the wide sense. The representatives are equivalent in the narrow sense if the new bases are defined over the same index sets as the old ones. From 8. we obtain (2) Every column-finite matrix is equivalent in the wide sense to a matrix 35, and in the narrow sense to a matrix T). 10. The theory of equivalence. It is now easy to give a complete account of the different equivalence classes, both for mappings and for matrices. From 9.2 and the connection between the equivalence of matrices in the narrow sense and the equivalence of mappings there follows (1) Let A be a linear mapping of E into F, let {xv}, veN, be a basis of E and let {y^}, fieN\, be a basis of F. Let A have rank r, nullity s and defect s'. If {xv<}, v'eN', is an arbitrary subset of r elements of the basis, whose complement {xv»}, v"eN"cN, has s elements, and if {yM'(v')}> /i'(v')eM'c M, is a subset of {y^} which is mapped in a one-one way, by fi'(v'), onto N', and whose complement {y^>}, /*"eM" cz M, has s' elements, then A is equivalent to the mapping given by (2) 0*v' = >V<v'), Dxv„=o (v'eN',v"eN") {xv>} is a basis of an inverse-image space of D, {jv(V')} a basis of the image space, {xv„} a basis of the null-space, and {y^} a basis of a complement of the image space. (3) Two linear mappings of E into F are equivalent if and only if they have equal ranks, equal nullities and equal defects. The sufficiency of the condition follows from (1). On the other hand, suppose that A2 = CA1B. If Alx = o, then CA1B(B~lx) = o and conversely. The kernel and inverse image space of Ax are therefore transformed into the kernel and inverse image space of A2 respectively,
1. The dual space 69 under the isomorphism B~l. Thus their dimensions remain unaltered. Finally the image space of A2 is equal to C(A1(E)), and so has the same dimension and co-dimension as AX(E). The three invariants of the equivalence classes cannot be chosen arbitrarily, since they clearly must satisfy the relations (4) r + s = d, r + s' = e, where d and e are the dimensions of E and F respectively. If d and e are finite, then it is clear that an equivalence class is determined by specifying the rank alone. For every triple of cardinal numbers r, s, s' satisfying (4), there is a mapping with these as invariants. If the rank, nullity and defect of a column-finite matrix 21 are defined as the rank, nullity and defect of the corresponding linear mapping (using 5.(3)), then the agreement of r,s and s' is necessary and sufficient for the equivalence of two matrices, in both the narrow and the wide senses. § 9. The algebraic dual space. Tensor products 1. The dual space. In §7,8. we became acquainted with two methods of obtaining new vector spaces from given ones. The most important construction of this sort however is that of the dual space, and we now turn to the study of this. Let E be a vector space over the field K. If an element u(x) = ux of K corresponds to each xeE, and if this correspondence is linear, i.e. if u^x^oLi + x2a2) = u(x1)al + u(x2)oi2 ^or a^ ^i?^2e^ and all aha2eK, then ux is called a linear functional or linear form on E. If we set (ul -\-u2)x = ulx + u2x and (olu)x = ol(ux) for aeK, then the set of all linear functionals on E clearly forms a vector space over K, the algebraic dual or algebraic conjugate £* of E. The prefix "algebraic" will be important later, but we shall leave it out for the time being. If £t and E2 are isomorphic, with xx<r+x2, then an isomorphism u{+-+u2 between Ef and #* ls clearly produced by setting u2x2 = uix1. (1) Isomorphic vector spaces have isomorphic dual spaces. It is easy to give an account of all the linear functionals on E. Let E have dimension d, and let {xv}, veN, be a basis of E. ux is known if all the values uxv = vv are known, for then we have (2) ux = u(YjXv^j =£(wxv)£v = £t;v£v.
70 § 9. The algebraic dual space. Tensor products On the other hand a linear form (2) is uniquely determined by an arbitrary choice of the values uxv = vv, for all veN. If to each ue£* we make correspond the vector u = {vv} then £* is mapped isomorphically onto the space cod(K) which was introduced in §7,8., and so onto the product of d copies of K with itself. If, as in §7,5., we replace x = £xv£v by the co-ordinate vector V * = {£v} Which corresponds to it under the isomorphism with cpd(K), then, by (2), ux is equal to the scalar product ux = ]Ti;v£v of the two vectors u and x; this is meaningful since only finitely many £v are nonzero. Combining these ideas together, (3) // E is a vector space of dimension d over K, then the choice of a basis xv of E leads to isomorphisms x<-+x and w<->u of E with (pd(K) and of E* with cod(K) respectively, under which ux = ux. In particular, (cpd(K))*^cod(K). It is easily verified that (4) If E=®Ea, then E*^T\E*. a a The dual space of a finite dimensional space E is isomorphic to E. The unit vectors evecod(K) are the images under the isomorphism (3) of the linear functionals wve£* defined by (5) uvxv=l, uvxv> = 0 for v' + v, v,v'eN. They form the dual system {uv} in £*, dual to the basis {xv} of E. 2. Orthogonality. We have just seen that a linear functional u on E is determined by its values on a basis of E. Remembering that a basis of a linear subspace can always be extended to a basis of the whole of E by adding further elements (§7,3.(2)), we immediately obtain the extension theorem (1) //' a linear functional l(y), yeF, is defined on a linear subspace F of a vector space E, then it can be extended linearly to the whole of E, and so there exists u0eE* with u0y = l(y) for all yeF. In a somewhat different version, we have (1') If x0 is an element of E which does not lie in a linear subspace F of E, then there is a u0eE* with uoy = 0 for all yeF, and with u0x0 = l. If ux = 0 for xeE and we£*, we say that u and x are orthogonal. If M is a subset of E, the set of all we£* which are orthogonal to all xeM forms a linear subspace of E*9 which we call the orthogonal space M1 of M in E*. If we start from a subset M of E*, we obtain the orthogonal space M1 of M in E in the same way. It follows directly from the definition that
2. Orthogonality 71 (2) // Mx c M2, then M£ <= M|, and so M^ c M^1. (3) M is always contained in M11. If M = M11, M is said to be orthogonally closed. An orthogonally closed subset of E or F* must be a linear subspace. M11 is called the orthogonal closure ofM. (4) M1 is always orthogonally closed. For by (3), M1c= M111; on the other hand, applying (2) to (3), we obtain M111 <= M1, and so M-l^M111. (5) Ffery /mear subspace F of E is orthogonally closed. For if x0 is not in F, then by (T) there is a i^eF1 with w0x0 = 1> so that x0 does not lie in F11; hence F = F11. On the other hand we have (6) //' E is infinite-dimensional, there are always linear subspaces, and indeed hyperplanes in F*, which are not orthogonally closed. We can take E = cpd(K), with dg:K0. The dual space cod(K) has F = <pd(K) as a proper linear subspace, by §7,8. But F1=o, for if x = {£v}eF1, then evs = £v = 0 for all v. Since F11 = cod(K), F is a proper subspace of F11. Every proper subspace lying between F and cod{K) is likewise not orthogonally closed, and in particular by §7,7.(7) there is a hyperplane which is not orthogonally closed. Nevertheless, we have (7) Every finite-dimensional linear subspace F of F* is orthogonally closed. This can also be expressed as follows: (7a) // u1,...,un are n linearly independent linear forms on E, and if F1 is the space orthogonal to F=[u1,..., un~], then each linear form on E which vanishes on F1 is a linear combination of ul9...,u„. F1 has codimension n in E. Proof. There is an xx with u1x1 = l. Suppose that it has been shown for k — \ that there are elements xl,...,xk_l in E such that UiXi=l and utXj = 0 for ij=l,...,k—l, and /=)=/. Then for each xeE we have k-1 (8) x= Yj xf(Mfx) + x;, with w,-x' = 0 for i= 1,..., k— 1. If ukx' = 0 for all x', then by (8) fc-1 x = 0, for all xeF, which is impossible, for the linear functional in the square bracket would then vanish identically. Thus there exists an xk with ukxk=l,
72 § 9. The algebraic dual space. Tensor products utxk = 0 for i=l,..., fc— 1. Putting xf = xf —(wfcxf)xfc, i = 1,..., k — 1, we have Mfxf=l, mi-xj- = 0 for j + i and ij = 1,..., k — 1; also ukxt = 0. If we now write xl9...,xfc for x1?..., xfc_l9 xfc, we have established the assertion for k. We have thus shown, for k = n, that E = F1@G, where G is the rc-dimensional subspace of E with basis x1?..., xn. If now u is a linear functional which vanishes on F1, and if ux( = vh then w — £ ufwf vanishes on all xeF, and so w = £ ufMf. i = 1 3. The lattice of orthogonally closed subspaces of £*. The orthogonally closed subspaces H of F* form a lattice V(E*) under the relation c=; the greatest lower bound of a set {Ha} exists in K(F*), being the intersection f] Ha. This is orthogonally closed, for using 2(2) it follows from H Ha cz ifa that (f) ^a)11 ^ H.1^^ so that (f) ^a)11 <= 0 ^- and the assertion follows from 2.(3). Similarly \J Ha exists as the a intersection of all the orthogonally closed subspaces which contain all Ha\ \J Ha is the orthogonal closure of the linear span of the Ha. a If to each linear subspace F of E we make correspond its orthogonal space F1, then by 2.(5) we have a one-one correspondence of V(E) with V(E*); by 2.(2) it transposes the partial order, so that, using § 7, 9. and 10., we have (1) The correspondence which sends each linear subspace F of E to its orthogonal space F1 cz F* is a dual isomorphism of the complete lattices V(E) and V{E*). V(E*) is thus also a complemented modular lattice. The lattice theoretic intersection in V(E) is the same as in V(E*), namely the set theoretic intersection. The union in V(E) is the linear span, and in V{E*) is the orthogonal closure. If F is infinite-dimensional, with basis xv, and if uv is the dual system in F*, the linear span of the one-dimensional and, by 2.(7), orthogonally closed spaces [wv] is different from the orthogonal closure F* of the [wv], by 2.(6). Since V(E*) is modular, however, it follows that (2) // Bx and B2 are orthogonally closed subspaces of F*, then Bx + B2 is always orthogonally closed, so that Blv B2 = Bl+B2. Suppose first that Bx is finite-dimensional, that u0eB1vB2 and that u0$B1+B2. Then the finite dimensional space C = Bi®\u0\ lies in V{E*). Since Bi cz C and V(E*) is modular, (3) (Bl vB2)nC = Bl v(B2nC).
4. The adjoint mapping 73 If veB2nC, then, being an element of C, it has the form v = w + u0ai, with weB1, aeK. But then u0ol = v — weB1+B29 which is only possible if a = 0. Thus v = weBx, B2nC cz Bl9 and 51v(52nC) = B1, so that, by (3), C = (B1v B2)nC = Bl, which contradicts the assumption about C. For general Br it follows from this that C lies in K(F*); (2) follows by repeating the argument. In particular it follows from (2) that (4) Two orthogonally closed subspaces of F* which are complements in the lattice sense are algebraic complements, and conversely. For, by (2), B1vB2 = E means simply that B1-\-B2 = E; together with B1nB2 = 0, this establishes the assertion. From (1), (4), and the fact that every linear subspace has a complementary space, there follows (5) If E = F1@F2 then E* = Fi®Fi If E* = Gl®G2 where G, and G2 are orthogonally closed, then E = G\®G2. Every orthogonally closed subspace of F* has an orthogonally closed complement. (6) // F is a linear subspace of F, then F1 ^(F/F)*. For if ueF1, then u(x + z) = ux, for each zeF. If we therefore define u' x, for xeE/F, as the common value of ux for all the xex, then each ueF1 corresponds to a u'e(E/F)*, different u corresponding to different u'. This correspondence is clearly linear. If, conversely, u'e{E/F)* is given, the formula ux = u'x defines a linear functional on E which lies in F1. Since u' is again obtained from this u by the first part of the argument, the correspondence is an isomorphism. In particular we have (7) // F has finite codimension n in E, F1 has dimension n. The assertion for orthogonally closed subspaces of F* corresponding to (6) is not in general true; we shall continue the investigation of the relationship between V(E) and K(F*), using other techniques, in § 10. 4. The adjoint mapping. Suppose that A maps F linearly into F. If v is a linear functional on F, then v(Ax), considered as a function of xeE, is a linear functional u on F, since i)[/l(x1a1 + x2fl(2)] = i;[(/lx1)fl(1 -\-(Ax2)a2]=v(Ax1)x1 +v(Ax2)a2. The mapping v^>u from F* into F* which is defined in this way is called the adjoint mapping A' of A. It is therefore defined by (1) (A'v)x = v(Ax), for all xeE, veF*.
74 § 9. The algebraic dual space. Tensor products A' is linear, since [A\(iivl+P2v2)]x = (li1v1+P2v2)(Ax) = pi[vMx)] + [i2[v2(Ax^ = ^[_{A'v1)x\ + ^2\_{A'v2)x\ so that we have (2) If Ae<5{E,F), then A'e<5(F*,E*). Let xv, veN, and yM, /ieM, be bases of E and F respectively. A is represented by a matrix 2l = ((aAiV)) with respect to these bases; with respect to the dual systems of these bases, v and u = A'v are represented by coordinate vectors v = {q>ll} and u = {vv} respectively. By 1. and § 8, 5., we have (3) vv = {A' v)xv = v{Axv) = v{S&tv) = YJ(p^^ This implies that u = 2Tt>, where 91' is the transposed matrix of 91; thus we have (4) // A is represented by the matrix 91 with respect to certain bases of E and F, then the adjoint mapping A' is represented by the transposed matrix 9T, with respect to the dual systems of the same bases. Since 91 is column-finite, 9F is row-finite. (5) {olA)' = olA9 {A + By = A+B'. for general A, B = ®(F,F), aeK. (BA)' = A'B\ for general Ae<Z(E,F\ £e6(F,G). We shall only establish the last relation: It follows from (1) that, for xeE, veG*, ({BAyv)x = v({BA)x) = v(B{Ax)) = {B'v){Ax) = (A'{B'v))x = ({A,B')v)x. The correspondence A^>A' defined by (1) is one-one, by (4), and so (5) follows from § 8, 6. (6) The algebra ®'(F) of mappings A' adjoint to the mappings A e ® (F) forms a sub-algebra of ®(F*) anti-isomorphic to ®(F). (7) // F is infinite-dimensional, ®'(F) is a proper sub-algebra of ®(F*). Proof. By (3), a mapping u = 9f o of cod(K) into itself is identically zero if the individual columns of 9F, which are the images 9F eM of the vectors eM, are all equal to the zero coordinate vector o. But if d^K0, there are linear mappings of cod(K) into itself which map all the eM into 0, but which map the vector e, all of whose coordinates are equal to one, and which is linearly independent of the eM, into a non-zero vector. 5. The dimension of F*. Because of the equation (1) (ax ux +ol2u2)x = ol1(u1 x)-\-a2(u2x)
5. The dimension of E* 75 each xeE defines a linear functional x(u) on E*, and so the correspondence x->x(w) is an isomorphism of E with a subspace of £**, which we can identify with E. If E is finite-dimensional, then £ = £**; on the other hand we have (2) // E is infinite-dimensional, then E is a proper subspace of £**, and so every term of the sequence E a £** cz £**** c ••• is a proper subspace of its successor. Proof. If £ is taken in the form cpd{K), then £* has the form cod(K). Since the vectors of q>d{K) are finite, they only define linear functionals which vanish on all but finitely many evecod(K). There exist linear functionals on cod(K), however, which are different from zero on infinitely many ev (use a suitable basis of cod(K), and proceed as in 1.(2)). The following theorem of Erdos and Kaplansky determines the dimension of £* exactly. (3) The dimension d* of cod(K) is, for infinite d, equal to the number kd of elements of cod(K), where k is the number of elements in K. Proof (W. Neumer). Let {uM} be a basis of cod(K), so that each uecod(K) is a linear combination of the vectors uM. There are (k— l)d* elements in cod(K) of the form uMa, aeK, a 4=0, [(/c — l)d*]2 elements of the form n^ocl +uM2a2, with ol1 and a2 both different from zero, and so 00 on, so that cod(K) contains £ [(/c-l)d*]' elements. Since p2=p for i = 0 oo oo every infinite cardinal p, we can write the sum as d* £ (k — 1 )l. £ (k — 1 )l i = 0 i = 0 is equal to X0, for finite /c^2, and to k, for infinite k; in either case we 00 therefore obtain d* £ (/c— iy = d*/c. This establishes the equation (4) I = ° /cd = d*/c, which we use to determine d*. If /crgd*, then d*/c:gd*2 = d*, so that d*/c = d*; the assertion d* = kd therefore follows from (4). We now show that /c:gd*, for all d^K0. It is clearly sufficient to prove this for d = K0. Let us therefore assume that the number d* of elements uM = (z;^, z;^,...) in the basis of co(K) is less than k. The collection M of all the coordinates uf of all the vectors uM has cardinal at most tf0d* = d*. Let K0 be the subfield of K generated by all the uf. Since K0 consists of rational expressions in the vf with coefficients in the prime field of K, KQ also has not more than K0d* = d* elements. Since d* <fc, there exists a ^ in K~K0. We set K1 = K0(^1), where K0(^) is the field obtained by adjoining ^ to K0. Again K{ has at most K0d* = d* elements. ^2GK~K1 can now be chosen, and the procedure repeated.
76 § 9. The algebraic dual space. Tensor products In this way we obtain a sequence K0c=...c=KMc=..., with K^K^CO, ^eK^K„_, Now let x = (£i) be the vector in co(K) formed by the £f, and let m x= Ysnfiir1i ^e *ts representation in terms of the basis. Because of the i = l linear independence of the vectors uM£, there exists an m x m sub-matrix ((i$)) of the matrix ({vf)), i = 1,..., m, j = 1,2,..., which has a non-zero determinant. The system of equations m can therefore be solved to determine the rjh and so the r\{ lie in K„o, for some sufficiently large n0. But it then follows from the fact that m £r= £ ujf^., for r>n0, that all the £r lie in KWo, which is impossible. i = 1 In particular, since (2Ko)d = 2d for d^K0, we have (5) // E is a real or complex vector space of dimension d^K0, then £* has dimension 2d. A detailed account of the relation between (pd(K{) and <pd(K2), where Kt is a sub-field of K2, can be found in Bourbaki [4], vol. 2, § 5. 6. The tensor product of vector spaces. Let E and F be two vector spaces over K. We form the set A(ExF) of all formal finite linear combinations ]T (x,y)axy of elements of ExF, with coefficients (x,y)eExF in K. A(£ x F) becomes a vector space over K, when we put [Z (*> J>) a*. y] P = Z (x> J>) a*, y /? and The zero element is obtained when all the coefficients ax y are put equal to 0. We write (x,j/) for (x,j;)l. We now form the linear span A0 in A(ExF) of all elements of the form (n m \ n m Zxt*i> Z ykPk)- Z Z (xf^k)a^fc- i = 1 fc = 1 / i = 1 k = 1 The quotient space A/A0 is called the tensor product, or direct product, E®F of £ and F. The coset to which {x,y)a belongs is denoted by (x(x)j/)a. Again, we write x®y instead of (x®y)l.
6. The tensor product of vector spaces 77 The following rule of calculation in E®F follows from (1): n m n m (2) £ x,ai® X yjk= X X (*«$)>*)«,&• i = 1 k = 1 i = 1 k = 1 Accordingly, o(g)j; = (oO)(g)j/ = (o(g)j;)0=o; likewise x®o=o. (3) The tensor product is commutative; that is, E®F is isomorphic to F ® E under the correspondence x ® y<r->y ® x. Since the mapping (x,y)-+(y,x) of E x F onto FxE defines an isomorphic mapping of /\(ExF) onto A(F x E) under which the subspaces A0 correspond, the quotient spaces are also isomorphic. It follows immediately from the equation (x® y)oc = x®(y<x) that n every element of E ® F can be written in the form ]T (xf ® yt). i= 1 (4) // yi,---,yM are linearly independent elements of F, it follows n from Yj (xi®yd =° tnat xi=° for i = 1,• • •,w. Proof. If u0eF*, a linear mapping A is defined from A(F x F) into F by putting A^(x,y)aix,y) = Y,y(uox)ax,r Since the elements (1) are mapped intoo, A induces a linear mapping A of E®F into F. Under this, £(xf®j/f) is mapped to ^^("o^)- If the j/f are linearly independent, it follows that £(xf(g)j;f)=o only if WqX—0 for all w0eF*, i.e. only if Xl.=o, i=l,...,n. (5) // {xv}, veN, is a basis /or F and {y^}, /ieM, is a basis for F, t/ierc Kg^h/ijeNxM, is a basis for E®F. If E has dimension d and F dimension e, then E® F has dimension de. For if X(xv®yM)aVM=o, then £(£xvaVM)<g) j;m=o, and so, by (4), V,fi fi V Yjxvavn=°> tnus avn = ^ f°r eacn (v,//)eNxM. On the other hand, V because of (2), every element of F ® F can be written as a finite linear combination of the xv®yfl. (6) // F has dimension d then E®F is isomorphic to the direct sum of d vector spaces isomorphic to F. For fixed fi, the terms xv®y^ form the basis of a subspace FM, by (5), and this is isomorphic to F under the correspondence (£xvav)(x));M ->£xvav; on the other hand, E® F = © FM. It follows easily from (5) that (7) // A and B are linear subspaces of E and F respectively, then A®B is isomorphic to the linear subspace of E®F spanned by the elements a®b, aeA,beB.
78 § 9. The algebraic dual space. Tensor products r (8) Every element z=ho of E®F has a representation z =r£x(l)' ®y{l) for which both the x{l) and the y{l) are linearly independent. i s t Proof. By (5), z = £ £ (*Vk® y^&ky ^ we introduce new bases fc=l 7=1 in £5 = [xvl,...,xvJ and i? = [yM1,...,yMJ, given by xVk = £x;/?,k and 3V,= Z y'm 7mj respectively, then m / m fc j so that the coefficient matrix SH = ((<xkj)) is replaced by $l = 932I(r, with 93 = ((/?/fc)), (£rrr((ymj)). Since arbitrary invertible matrices can be chosen for 93 and (£, if they are chosen suitably, the matrix & has the equivalent canonical form (cf. §8,9.(2)). If r is the rank of 21, then r z = Y^x'i ® y'i is tne required representation. i Calling the number r which appears in (8) the rank of z, we have (9) The rank of an element zeE®F is equal to the rank of the matrix ((av/J), for any representation z = £X(xv(x) J^)0^ in terms of bases {xv} and {y^} of E and F. v v When we introduce new bases by xv = £xv,/?v,v and yM==X jyy^, v' \i' only finitely many non-zero av>, appear in the expression z = £(xV®Jv)(*v>' ^ follows that we can obtain this representation of z from the first by a change of basis within finite-dimensional sub- spaces of E and F. For this, as in the proof of (8), the rank of the coefficient matrix remains unaltered, and in particular it is equal to the rank of z. 7. Linear mappings of tensor products. A mapping B(x,y) from Ex F into a vector space H which is linear in both variables is called a bilinear mapping from ExF into H. Thus, for all xteE, ykeF, (1) B(Zxiai,Ylykpk) = YlYlB(xt,yl)alpk. Vi k J i k If H is the field K of coefficients, we speak of bilinear forms, or bilinear functionals. The set of all bilinear mappings from ExF into H forms a vector space %>(ExF,H). We denote the space of all bilinear functionals on ExF by ^B(ExF). (2) Every bilinear mapping B of Ex F into H defines a linear mapping B of E®F into H, and conversely. This correspondence is an isomorphism between $5(Ex F,H) and <5(E® F,H).
7. Linear mappings of tensor products 79 Putting B[ Yj (x>y)<*x,y\= Z B(x,y)ax y, B ^ extended linearly to Ux,y) ' J (x,y) the whole of Aj^FxF). B vanishes on A0, by (1), and so defines a linear mapping B of E®F into H. If, conversely, B is a linear mapping of £®F into H, then we define a mapping £ from ExF into H by the equation B{x,y) = B(x®y); this is bilinear by 6.(2). Since B and B are determined by their values at the points (x,j/) and x®y respectively, the correspondence is an isomorphism between 93(FxF,H) and S(E<g)F,H). (3) // F, F and H are vector spaces, S(£(g)F,H)^S(£,S(F,H)). It is sufficient, by (2), to show that 93(F x F,ff)^S(E,S(F,ff)). If z = #(x,j/) is a bilinear mapping of ExF into H, for fixed x let £x be the linear mapping which sends each yeF to the element (4) Bxy = B(x,y) = zeH. In this way B corresponds to the linear mapping x->2Jx from E into S(F,H). Conversely, given a linear mapping x^Bx from F into S(F,H), a bilinear mapping is defined from F x F into H, using (4) again. This correspondence is clearly one-one and linear. If H=K, we obtain as a special case (5) The vector space (E®F)* is isomorphic to &(E,F*), and to 93(FxF). How is F* (x) F* related to (F <g) F)*? (6) F*(x)F* can a/ways be interpreted as a subspace of (F(x)F)*. The two spaces coincide if and only if E or F is finite-dimensional. Proof. If, for weF*, ueF*, xeF, yeF, we put (u,v)(x,y) = (ux){vy), we obtain bilinear forms on F x F and on E^xF*, which, by (2), determine linear forms (7) {u®v)(x®y) = (ux)(vy), on E® F and F* (x) F* respectively. In this way every element of £* (g /?* js interpreted as a linear functional on E®F. Different elements of E* ® F* determine different linear functionals: for this, it is enough to show that an element ]T u{ ® vt with linearly independent vt and non-zero ut does not vanish identically. n The linear functional on E®F determined by £ ut®vt corresponds, by (4) and (5), to the linear mapping I = 1 n (8) x-^i^x)^.,
80 § 9. The algebraic dual space. Tensor products from F into F*. This does not vanish identically, because the vt are linearly independent. The mappings (8) in <S(E,F*) all have finite-dimensional image spaces in F*. But if E and F are both infinite-dimensional, there are mappings in S(E,F) with infinite-dimensional image spaces, so that in this case E* ® F* is a proper subspace of (F(x)F)*. Finally we show that F* ® F* = (E ® F)* if F is finite-dimensional. Let {xf} and {j/M} be bases of F and F respectively, so that each we(E®F)* is determined by its values w(xf(g) 3/^ = 1/^. For each i, there is a u(0eF*, with v{i)y^ = il/ifi for all /i. If {wj is the basis dual to {xj, then I ^ w7(x) ^0) J (x^ (x) yM) = (//t> for all i, \i, so that w= Yj Uj®vU)eE* ® F*. Similarly, we show that, provided F,F,G,//=ho, (9) S(F,F)®S(G,i/) can a/ways be interpreted as a subspace of <Z(E®G,F®H). Equality holds if and only if either E and F, or G and H, or E and G are finite-dimensional. ((6) is a special case of (9), obtained by replacing F and H by K, and observing that K® K^K, by 6.(6).) Proof. As above, a linear mapping (10) (A®B)(x®z) = (Ax)®(Bz)9 from E®G into F®H is defined by putting (A,B)(x,z) = (Ax)®(Bz) for Ae<5{E,F), Be<5{G,H), xeE and zeG, and using (2). To the n mapping £ At <g) B; there corresponds, by (4), the mapping (11) x-»XUiX)®B, I = 1 from F into S(G,F®H). If the Bf are linearly independent, then (11) is different from zero, and so we obtain the representation of S(£,F)(g)S(G,H) as a subspace of <Z(E®G,F®H). As x runs through all of F, we obtain on the right-hand side of (11) only those mappings from G into F®H which comprise the same Bt. A mapping ]T A{ ® B{ thus can only correspond to a mapping of F into S(G,F®H) which sends F into a linear subspace of mappings contained in the linear span of mappings of the form y®Bh where yeF, i=l,...,n. If G or H is infinite-dimensional, there are infinitely many linearly independent Bae&(G,H). If F is also of infinite dimension, then there is a linear mapping of F into ^(G^F® H) whose image
7. Linear mappings of tensor products 81 space contains infinitely many y0 ® Ba, where y0 is a fixed non-zero element of F. This mapping cannot be produced by a ]T A(®Bi. If E and at least one of G and H are infinite-dimensional S(E,F)® S(G,J/) is therefore a proper subspace of &(E®G,F®H). Since £ and G are on an equal footing, it also follows that if G is infinite- dimensional, equality is only possible if E and F are finite-dimensional. The conditions of (9) are therefore necessary. It remains to show that the relation (12) S(£,F)®S(G,H) = S(£®G,F®H) holds in the specified cases. a) Let E and G be finite-dimensional, with bases {xn} and {zk}. If {y^} and {tx} are bases of F and H, a mapping C from E®G into F®H is determined by C(xw® zJ = £]TywfcM;i(j/M® tA); only finitely many y are different from zero. If Ajtl is the linear mapping in S(E,F) which sends .Xj to y^ and x- to o, when j+f, and if, correspondingly, Bue<Z(G,H) sends z, to tk and zr to o, when /#=/', then (Z Z Z Z Vjima^jm ® Bii) (x« ® zfc) = Z Z twaOv ® '*)> so that (12) holds. b) Let E and F be finite-dimensional, with bases {xn} and {ym}, and let {zK} and {t^} be bases for G and H respectively. A mapping CeS(£®G,F®#) is determined by C(xn® zK) = YJYjynKmX(ym®tj. m A Let X„me6(£,F) be defined by i4IImxII = j;lll, and AMmxM,=o, for n^n'; let Bnme(5(G,H) be defined by £„wzK = £ 7™™a'a, f°r all k. Then ( Z ^'m ® Bn'n) (** <g> zj = £ £ ^mA^m ® **), n'm ' m A so that again (12) holds. c) The case where G and H are finite-dimensional follows from b) by interchanging E and F with G and H. If A and £ in (10) have the form H V A K with respect to bases {xv}, {j/M}, {zK} and {t;}, then there are corresponding matrices 9I = ((aMj) and 93 = ((/JAk)). With respect to the bases xv®zK and yM®rA, the mapping ,4®# is represented by the matrix ((aMV^;J), which is called the Kronecker product 91®95. This clearly results from replacing the element aMV of the matrix 91 by the matrix aMV95. 6 Kothe, Topological Vector Spaces I
82 § 10. Linearly topologized spaces § 10. Linearly topologized spaces 1. Preliminary remarks. There is an evident symmetry between a vector space E over a field K and its dual space £*, as our earlier considerations have shown. If E is finite-dimensional the symmetry is complete, as E and £* are isomorphic, and each can be interpreted as the dual of the other. If E is infinite-dimensional, however, E does not comprise-all the linear functional on E*. In the same way, the correspondence which sends each linear subspace of E to its orthogonal space only produces the orthogonally closed subspaces of £*. Finally S(£) is anti-isomorphic to the algebra <S'{E) of adjoint mappings, and S'(E) is a proper subalgebra of S(E*) if E is infinite-dimensional. It is natural to ask whether this complete duality between E and £* in the finite-dimensional case can be obtained in the general case as well, by introducing some concept of continuity, and so some suitable topology, in E and E*. This must happen in such a way that the continuous linear functional on £* must be just those which are determined by the elements of E. We can use this fact to derive a condition on the topologies on £* under considerations. A linear functional l(u) on £* maps £* into the field K. In order to be able to speak about a continuous linear functional, a topology must be defined on K. Every Hausdorff topology on a field with finitely many elements is the discrete one, by §1,5.(2). In order to obtain a theory valid for general fields, throughout this chapter we shall suppose that K is discrete. Now if l(u) is to be continuous at o, there must be a neighbourhood U(o) for which /(w) = 0 for all ueU{6). Since l(u) is linear, the linear span of U satisfies this condition, if U does. The continuity of l(u) at a general point u0e£* is then satisfied if we take U(u0) to be equal to the linear manifold u0 + U(o). It will therefore be appropriate for our purpose if we restrict ourselves to topologies for which there is a base of neighbourhoods of o consisting of linear subspaces, and for which the neighbourhoods U(u0) are obtained by taking the translates w0+ U(o) of the neighbourhoods of zero. Following Lefschetz [1], we call such a topology linear. 2. Linearly topologized spaces. Let L be a vector space over K. Suppose that a topology X0 is defined on K. A topology X defined on L is said to be compatible with the vector space operations in L if ax and x + y are jointly continuous in both variables. For x + y this means that the mapping (x,j/)->x + j/ from LxL into Lis continuous, when LxL is given the product topology. Similarly (x,x)->Ax must be a continuous function on the topological product K x L, with values
2. Linearly topologized spaces 83 in L. If K is discrete, it is sufficient to establish the partial continuity of ax on L. If 3: is Hausdorff and compatible with the vector space operations, then L is called a topological vector space over K. For brevity, we also write this as L[2T|. A topological vector space over a discrete field, with a topology which is linear in the sense of 1., is called a linearly topologized space. We now establish (1) Let {Ua}, aeA, be a filter base consisting of linear subspaces of the vector space L, with f] Ua=o. If we introduce the linear topo- a logy X on L defined by taking {Ua} as a base of neighbourhoods of o, then L is a linearly topologized space with respect to X. In addition, we show (2) The topology X is produced by a Hausdorff uniformity on L, so that, in particular, L is regular. We take for a base of the uniformity on L the sets Na of all (x,y)eLxL with x — yeUa. The Na form a filter base on LxL, since the Ua form a filter base on L. Because f] Ua=^)f>\^N>0i is the diagonal, a a i. e. the set of all (x,x) in LxL. Thus (V 1) and (V4) of § 5,1. and 2. are established. (V2') is trivial, and finally (V3) holds, since N* = Na, for each a. The uniformity defined in this way is therefore Hausdorff. The topology corresponding to it is X, and regularity follows from § 5, 2.(3). Finally X is compatible with the vector space operations: ax is continuous in x, since ?.xeUa if xeUa. If (x,y)e(x0 + UOL,y0 + UJ, then x + yex0 + y0+ Ua, and so x + y is also continuous in both variables. We shall also denote the neighbourhood x0+Ua by Ua(x0). (3) Every linear neighbourhood U of o is open and closed; more generally each set M +U, M c L, is open and closed. Every linearly topologized space is totally disconnected. M + U is open, since x+U lies in M+U if x does. But L~(M+U) is also open, since no point of y+ U lies in M + U if y does not belong to M + U. If the subset S of L contains two points x and y, and if U(x) is a linear neighbourhood of x which does not contain y, then SnU(x) is a proper open and closed subset of S. No subset of L with more than one point is connected, therefore, and the result follows from §1,6. The next result is immediately obvious: (4) A vector space L is a linearly topologized space, when it is given the discrete topology. (5) Every finite-dimensional linearly topologized space is discrete. 6-
84 § 10. Linearly topologized spaces For it follows from f] Ua=o that finitely many linear U have inter- a section o, and so o is a neighbourhood of o. (6) Let Lx be a linear subspace of L[Z~\. (a) The closure Lx is again a linear subspace. (b) Lx is a linearly topologized space under the subspace topology induced by X. Proof, (a) Let x0 and y0 be two closure points of L{. For each linear neighbourhood U of o there are elements x,yeL{ with xex0+U, yey0 + U; xa + ypex0<x + y0P+U for arbitrary a, /?eK, so that x0a + y0peLv (b) As Ua runs through a linear base of ^-neighbourhoods of o, the sets L{nUa are linear subspaces of L1 with intersection o, and they form a base of neighbourhoods of o for the induced topology. (7) The topological product TT Lp of linearly topologized spaces is again a linearly topologized space. This follows without difficulty from § 7, 8. and §1,8. If we define a topology X on the direct sum © LB of linearly topo- P logized spaces Lp, with topologies Zfi9 by taking as neighbourhoods of o the direct sums © UB, where each UB is a linear neighbourhood P of o in Lp, we again obtain a linearly topologized space, which we call the topological direct sum © LB\TB~\ of the Lp. Thus (8) The topological direct sum L[X] = @Lp[Xp] of linearly topologized spaces is again a linearly topologized space. We remark that the topology induced on Lp by % is the original topology Zp. Two linearly topologized spaces L^XJ and L2[£2] are said to be topologically isomorphic, and we write L1[^X1]^L2[^X2]5 if there is a linear one-one correspondence between L1 and L2, which is continuous in both directions. n n For finitely many Lf, TT Lf[3:f] and © L.-pJ are topologically isomorphic (§ 7,8.). / = 1 I = 1 (9) // a linear mapping A from Ll\Zl~\ into L2[^X2] is continuous at o, then it is continuous everywhere, and is also uniformly continuous. It is enough to establish uniform continuity. By (2), a basic vicinity N of L2 consists of all {yl,y2) with yl— y2eK, where V is a neighbourhood of zero in L2. By hypothesis there is a neighbourhood U <=-Lx with A(U)czV, and so, since Ax{ — Ax2 = A(xl — x2\ the image of the vicinity consisting of all those (xux2) with xl—x2eU is contained in V.
3. Dual pairs, weak topologies 85 3. Dual pairs, weak topologies. Dieudonne and Mackey have introduced an idea which has proved to be particularly fruitful in the study of topological vector spaces. Two vector spaces Lx and L2 over K form a dual pair or linear system (L2,Ll} when an element of K, denoted by ux or <u,x>, is associated with every pair (u,x)eL2x Lx, in such a way that the following hold: (D1) ux is a bilinear form, i. e. u(xl(xl +x2a2) = (uxl)fxl +(«x2)a2, (ftuj + P2u2)x = j^l(ul x) + fi2(u2x). (D2') If, for some xeLi9ux = 0 for all ueL2, then x=o. (D2") If, for some ueL2,ux=o for all xeLx, then u=o. By (Dl), each ueL2 determines a linear functional in L*l5 and distinct u determine distinct linear functional, by (D2"); thus L2 can be interpreted as a subspace of U[, and Z^ as a subspace of L*2. The conditions (D2) say that Lx and L2 contain "sufficiently many" linear functional in L*2 and L*u respectively. We say that two dual pairs {L2,L1} and (L2,L\) are isomorphic, and write (sL2,Liy = (^L2,Lls), if there are linear one-one correspondences x<->x and u<^>u between Lt and Lx and between L2 and L2, for which ux = ux always holds. In these terms, the results of §9, 1 can be formulated as follows: (1) A vector space E and its algebraic dual £* form a dual pair <£*,£>. If E has dimension d over K, <£*,£>^<cod(K),(pd(K)>. As in §9,2., we call the set M1 of all ueL2 with ux = 0 for all xeM czLx the space orthogonal to M in L2. M a Ll is said to be orthogonally closed with respect to L2 if the subspace M11 of L1 orthogonal to M1 is equal to M. Theorems § 9, 2.(2), (3) and (4), are still valid, and as in § 9. 3. imply (2) Let (L2,L{y be a dual pair. If we make each orthogonally closed subspace of Lx (respectively L2) correspond to its orthogonal space, then the correspondence is a dual isomorphism between the complete lattices V{LX) and V(L2). Given a dual pair {L2,L1}, we have the following natural construction for a linear topology on Lx (with a corresponding one for L2): we take as base of neighbourhoods of o all sets UUu ^Un, uteL2, consisting of all those xeL{ for which (3) u(x = 0, i= l,...,n. A neighbourhood determined by this base of neighbourhoods is called a weak neighbourhood of o. The weak neighbourhoods of o in Lx are just the spaces F1 orthogonal to the finite-dimensional subspaces F
86 § 10. Linearly topologized spaces of L2, and the subsets of L1 containing them. The topology defined on Lx by L2 in this way is called the linear weak topology on Lx with respect to L2, and is denoted by Zls(L2). In the same way Lx determines the weak topology %ls(Lx) on L2. (4) //' {L2,L1} is a dual pair, L1[_<£ls(L2)~] is a linearly topologized space. The intersection condition of 2.(1) is satisfied, because of (D2). 4. The dual space. Given a linearly topologized space L[3T|, we can consider the continuous linear functionals on it. If ux and u2 are continuous linear functionals on L[3f[, so are oluu aeK, and u{+u2, and so the collection of continuous linear functionals forms a vector space, which is called the dual space or conjugate space L[3T|' (or more simply L) of L[%~] with respect to X. The algebraic dual L* is nothing but the dual of L with respect to the discrete topology. Thus we always have L c L*. As in §9,2., the following extension theorem holds: (1) If a continuous linear functional /(y), yeF, is defined on a linear subspace F of the linearly topologized space L, then it can be extended continuously to the whole of L. Thus there exists ueU with uy = l(y), for yeF. (V) If x0 is an element of L not lying in the closed linear subspace F of L, there exists a u0eL with u0x0 = \ and uoy = 0 for all yeF. First we establish (V). By hypothesis, there exists a linear neighbourhood U of o for which (x0 + U) n F is empty. F + U is therefore a linear subspace which does not contain x0. By §9,2.(1'), there exists w0eL* with w0x0 = l and uoz = 0, for all zeF+U. Since u0 vanishes on the whole of U, u0 is continuous, and so belongs to L. It is sufficient to establish (1) for a linear functional l(y) which does not vanish identically on F. If l(y0) = 1, F can be written as F = \_y0~\ © Fx, where l(y0)=\ and l(z) = 0 for all zeFx (put y = ay0-^(y — ay0\ for arbitrary yeF, where l(y) = a). Since / is continuous on F, there is a neighbourhood U of o in L, for which the induced neighbourhood (j/0 +U)nF in F has an empty intersection with Ft. But it follows from this that y0 does not lie in Ft + U, and this is closed, by 2.(3); applying (T), we obtain a u0eL which vanishes on Fx + U, and which takes the value u0y0 = \. Thus u0 coincides with l(y) on F. (2) Every linearly topologized space L, together with its dual space L', forms a dual pair (L,L}. Since L <= L*, conditions (D1) and (D2") are satisfied. (D2') follows from (1), for if x + 0, there is a non-zero continuous linear functional on the one-dimensional subspace [x] <= L, by 2.(4), (5) and (6 b); using (1), this can be extended to ueL\ with wx + 0.
4. The dual space 87 (3) // (L2,Lxy is a dual pair, then LX=L2 and L2 = LU the duals being taken with respect to the linear weak topologies. Proof. It is enough to show that L\=L2. (a) Every ueL2 defines a weakly continuous linear functional on Lu since ux = 0 for each xeUu, and so u is continuous at o. (b) If, conversely, u is a weakly continuous linear functional on Lu there is a weak neighbourhood F1 = [u1,...,un]1 of o, with the ut independent, on which u vanishes. By § 9, 2. (7 a), u is a linear combination of the ui9 and so lies in L2. (L,L} forms a dual pair, by (2), and so, using 3., the weak topology Zls(L) can be introduced on L. Its relation to the original topology X is given by (4) // L[X] is a linearly topologized space, *% is always finer than the linear weak topology £/s(L'). If u0eL, there is a ^-neighbourhood U of o with uox = 0, for all xeU. Every hyperplane uox = 0 is thus a ^-neighbourhood of o. Every basic ^-neighbourhood of o is therefore a ^-neighbourhood of o, as it is the intersection of finitely many such hyperplanes. (4) can also be expressed as (5) Zls(L!) is the coarsest linear topology on L for which L is the dual space. (6) // F is a X-closed linear subspace of L[X], F is orthogonally closed with respect to L; conversely a linear subspace F of L which is orthogonally closed with respect to L is %ls(L)-closed. In particular the H-closed and Xls(L)-closed linear subspaces of L[£] coincide. The closure F of an arbitrary linear subspace F cz L is equal to F11. Proof. (1') implies that a ^-closed subspace is orthogonally closed. Suppose conversely that F11 = F, and that x0 is a weak closure point of F. If u0 is any point of F1, there is at least one yeF in the weak neighbourhood UUo(x0), and so u0x0 = u0(x0 — j/) = 0. Since this holds for each u0eFL, x0 lies in F1JL = F. The last assertion now follows from 2. (6 a). (7) // F is a closed linear subspace of the linearly topologized space L, every linear subspace G in which F has finite codimension is also closed. It is enough, by (6), to show that G = F © [x0] is orthogonally closed. By (T) there exists u0eFL with w0x0 = l. If u is an arbitrary element of F1, v = u — (ux0)u0 belongs to G1. If zeG11, we must have vz = uz — (uxo)uoz = 0, and so u(z — x0(u0z)) = 0, for all ueF1. This means that z — (u0z)x0eF, that zeG, and that G1L = G.
88 § 10. Linearly topologized spaces (8) The linear weak topology of a linearly topologized space L[X] has as base of neighbourhoods of o the collection of all X-closed linear subspaces of finite co-dimension. If F is an n-dimensional linear subspace of L', F1 is of codimension n in L, by § 9, 2. (7 a). Conversely, a X-closed linear subspace of codimension n is intersection of n hyperplanes Hh by § 7,7.(7), and these are ^-closed, by 7. H{ is the space orthogonal to a one-dimensional subspace [w-] c L', by 4.(1'), and so H = [w1,...,wII]1. Another proof of (8) will be given in 9.(5). 5. The dual pairs <£*,£>. We now inquire whether we can answer the problem in 1., using the results which we have established. We give E and F* the linear weak topologies %ls(E*) and Zls(E) respectively. Then, by 4.(3), E* = E' and £ = (£*)', so that the linear functional on F* which lie in E are indeed characterized as the weakly continuous ones, whereas conversely all the linear functionals on E are weakly continuous. By 4.(6), the orthogonally closed subspaces of F* are precisely the weakly closed ones; on the other hand, by 4.(6) and §9, 2.(5), all the linear subspaces of E are weakly closed. Further, we have (1) // y4eS(£,F), the adjoint mapping A is a weakly continuous mapping from F* into F*; conversely every weakly continuous linear mapping from F* into F* is the adjoint of some /le£(F,F). (a) It is enough to consider continuity at o. Let Ae&(E,F). Given Uxl9...9Xn(o) in F*, the image A'v of any v in the neighbourhood U axx*'-">axS°) m ^* lies in Uxl,...,Xn(o\ since (A'v)xi = v(Axi); hence A is weakly continuous. (b) Conversely, let B be a weakly continuous linear mapping of F* into E*. For each x0eE, (Bv)x0 is a weakly continuous linear functional on F*. Since (F:¥)f = F, there is a uniquely determined y0eF for which (Bv)x0 = vy0. We put y0 = Ax0. The mapping defined in this way is clearly linear, and so XeS(F,F). It follows immediately from (Bv)x = v(Ax) that B = A. (2) Every Ae&(E,F) is weakly continuous. For given a neighbourhood UVl,...,Vn(o), v^eF*, in F, Ax lies in this neighbourhood if xeUA,Vl,...,A,Vn(o). Using § 9,4. (6) in the case where E = F, (1) and (2) imply that (3) The algebras L(E) and L(E*) of weakly continuous endomorphisms are anti-isomorphic.
6. Weak convergence and weak completeness 89 The weak topology therefore satisfies all the requirements made in 1. The next section shows, however, that it is not satisfactory in every respect. 6. Weak convergence and weak completeness. If E is infinite-dimensional, every basic weak neighbourhood of o is a linear subspace of E of finite co-dimension. The topology Zls(E*) on E is therefore different from the discrete topology. Nevertheless, the two topologies give the same criterion for the convergence of sequences, as we shall now show. A sequence xn cz E is said to be almost constant if all the terms of the sequence are the same, from a certain n0 onwards. A sequence is clearly a Cauchy sequence for the discrete topology if and only if it is almost constant. On the other hand, we also have (1) Every weak Cauchy sequence in E is almost constant. Xls induces the discrete topology on every finite dimensional sub- space of E. It is therefore enough to show that a weak Cauchy sequence lies in a finite-dimensional subspace of E. If this were not so, it would be possible to find a subsequence xn of linearly independent terms, and it would then be possible to define a linear functional u0 on E with u0xn.=\=u0xn._l9 since K has at least two elements. But there would then be no n0 beyond which the terms x — x„ , would lie in UUo(o). An immediate consequence is (2) E is weakly sequentially complete. In order to describe weak convergence in £*, we establish (3) // E has dimension d over K, £*, with the topology %S{E\ is topologically isomorphic to the topological product Kd, with K discrete. Let {xv}, veN, be a basis for E, and let {uv} be the dual system. Let [wv] be the one-dimensional subspace of £*, isomorphic to K, with basic element uv. By § 9,1., £* is isomorphic, as a vector space, to TT[wv]. Xls induces the discrete topology on [wv], and the sets UXvi f...iJCvn(o) form a base of neighbourhoods of o in E*. The result therefore follows from § 1, 8. From this there follows immediately (4) If {xv}, veN, is a basis of £, a sequence uneE* is a weak Cauchy sequence if and only if every sequence unxv, n= 1,2,..., is almost constant. In particular, a sequence u{n) = {v("}\ in cod(K) is a weak Cauchy sequence if and only if every sequence v[n\ w = l,2,..., is almost constant in K. It follows that £* is weakly sequentially complete. We can say even more, however. (5) £* is weakly complete.
90 § 10. Linearly topologized spaces For £* is topologically isomorphic to TT[wv], by (3). Each [wv] V is a discrete topological space, and is complete, by §5,4.(3). But then TT [wv] is complete, by § 5, 7.(2), and so, therefore, is £*. V In contrast, we have (6) // E is infinite dimensional, E is not weakly complete. Considering E as a subspace of £**, E** is the weak closure of E. Here £** is to be considered as a linearly topologized space under Xls(E*). £ is a proper subspace of £**, by § 9, 5.(2). The space E1 orthogonal to E in £* is o, so that the space E11 orthogonal to E1 in £** is £**. Hence £** is the orthogonal closure of E, and the result follows from 4. (6). The linear weak topology is therefore not completely satisfactory, since E and £* are not both weakly complete. It is reasonable to try to replace Xls by a finer linear topology, which coincides on E with the discrete topology, but which does not alter the results for E*. The solution of this problem will be given as we investigate linearly topologized spaces further; we turn to the study of these once more. 7. Quotient spaces and topological complements. Let L[X~] be a linearly topologized space, and let X be defined by linear neighbourhoods Ua of o. Let Ll be a linear subspace of L. By 2.(6), it is again a linearly topologized space. A topology can be introduced in a natural way on the quotient space L/Ll, by taking as open sets in LjLx the images O of the open sets O in L, under the canonical mapping K of L onto L/Ll (cf. § 8, 2.). If O is open in L, so is 0-\-Lx, and so K{~ 1}(0) is again open. From this it follows easily that (01) and (02) of § 1,1. are satisfied; for example, n f]6i is again open, since fl^n^'1^^^^"1^,-). i = 1 The topology defined in this way on LjLx is called the topology induced by X; it is again denoted by X. This topology can only be Hausdorff if Lx is closed in L, for if x0 is a closure point of Lt which does not lie in Ll9 then x^6 belongs to every neighbourhood of 6. On the other hand, we have (1) // Lx is a closed linear subspace of L\X\ L/Ll is a linearly topologized space under the induced topology X. Proof. The induced topology is linear, since the images \J% of the Ua are linear subspaces of L/L1, and they form a base of neighbourhoods ofo. Further, L/Lx is Hausdorff if there is for each x0=|=6 a neighbourhood ofo which does not contain x0. For if Ua is a linear neighbourhood
7. Quotient spaces and topological complements 91 contained in this neighbourhood, x0 + Ua and the neighbourhood Ua of 6 are disjoint. But the set of all x=£x0 is open and contains 6, since it is the K-image of the complement of the closed set x0 + Ll. (2) // Lx is a subspace of L which is both open and closed, then L/Lx is discrete. For Lx =6 is a neighbourhood of 6. A continuous linear mapping A from one linearly topologized space Lx into another L2 is called a topological homomorphism if it is open - that is, if every open set in Lx is mapped into an open set in A(LX). MA is also one-one A is called a topological monomorphism of Lx into L2. A is then a homeomorphism of Lx and A(LX) (cf. § 1, 7.). If A(Ll) = L2 as well, A is called a topological isomorphism of Lx and L2. It follows directly from the definition of the induced topology that (3) // Lx is closed, the canonical mapping K of L onto L/Lx is a topological homomorphism. (4) A continuous linear mapping A from Lx into L2 always has a closed null-space N[A~\. A is the product of the canonical homomorphism K of Lj onto Ll/N[A], a continuous one-one linear mapping A of LX/N[A~\ onto A(LX), and the embedding J of A(LX) into L2. Proof. The null-space N[A~] is closed, since it is the inverse image of the closed set {o} cz L2. By (3), K is therefore a topological homomorphism. The inverse image A(~l)(0) of an open set O cz A(LX) is open, and so therefore is K A(~ 1](0) = A{~ 1](0); thus A is continuous, and it is clearly one-one. Further, we have an analogue to §8,2.(1) (5) Every topological homomorphism A of Lx onto L2 is the product of the canonical homomorphism K of Lx onto Ll/N[A~\ and a topological monomorphism A of L1/N[A~\ into L2. The topological monomorphism A of L{/N\_A~\ into L2 is the product of a topological isomorphism A of L1/N[A~\ onto A(Lx), and the embedding J of A^L^) into L2, which is a topological monomorphism. Conversely, the continuous linear mapping A is a topological homomorphism if A is a topological isomorphism. The simple proof, which uses (4), can be omitted. Let F cz G be two closed linear subspaces of L[£], and let Kx be the canonical homomorphism of L onto L/F. The image K{(G) is G/F. If we apply the canonical homomorphism K2 of L/F onto (L/F)/(G/F) to L/F, K2Kl is a topological homomorphism of L onto (L/F)/(G/F),
92 § 10. Linearly topologized spaces with N\_K2Kl~] = G, and so by (5) we have a ^-isomorphism (L/F)/{G/F) = L/G (cf. §7,6(7)). The other of the two isomorphism theorems, F/F nG = (F + G)/G, is not generally true in the topological sense for closed linear subspaces F and G of L[£] (cf. § 13, 6.). Two complementary closed linear subspaces Lx and L2 of L are called ^-complementary if L[£] is the topological direct sum (the topological product) of ^[JX] and L2[3T], in the sense of 2. (6) A closed linear subspace Lx of L\H\ has a %-complement L2 if and only if there is a continuous projection P of L onto L1. Pl is then a topological homomorphism, L2=N\_P1~\, and L/L2 is topologically isomorphic to L1. Proof, a) If L[I] = L1[I]xL2[I], the projection Pt of L onto Lx is continuous, by § 1, 8. b) On the other hand, if Px is continuous, and if N[P1~] = L2, then L2 is closed, and by § 8.3 it is an algebraic complement of Lx. Moreover the projection I — P1=P2 of L onto L2 is continuous. Each x = xx+x2, xleLu x2eL2 corresponds in a one-one linear way to an element x = (xl,x2)eLl[X] x L2[X] = L. The mapping x-+x from L onto L is continuous, for given a neighbourhood U of o in L, it is sufficient to take xe([/nL1)©([/nL2). Conversely the mapping QiX = (x1,o) is continuous, as it is the product of the continuous projection Plx = x1 and the continuous mapping x1->(x1,o); the mapping (2i+Q2)x = x is therefore also continuous. c) The projection of L{ x L2 onto Lx is a homomorphism since it is open, by § 1, 8. The last assertion follows from (5). (7) Every algebraic complement of a linear neighbourhood U of o in a linearly topologized space L[X] is a discrete %-complement of U. If H is an algebraic complement of U, H nU=o is an open set in H, and so H is discrete. The projection of L onto H is continuous, for every subset M of H has an open inverse image M + U, by 2.(3). The assertion follows from (6). (8) Every finite dimensional linear subspace G of a linearly topologized space L[X] has a %-complement. By 2.(5), G is discrete. If xl,...,xn is a basis of G, then by 4.(1) there are elements ul9...9uneU9 with wfxf=l and w,xfc = 0, if i^k. Let F be the n-dimensional linear subspace of L spanned by the ut. F1 is a ^-neighbourhood of o, and L=FL®G9 by §9, 2. (7 a). F1 and G are ^-complementary, by (7). (9) Every linearly topologized space L[X] is topologically isomorphic to a linear subspace H of a topological product of discrete spaces.
8. Dual spaces of subspaces and quotient spaces 93 Let {Ua} be a base of ^-neighbourhoods of o consisting of linear subspaces. By 2.(3), each Ua is open and closed, and, by (2), L/Ua = La is discrete. We denote the coset of x with respect to Ua by xa. Let L be the topological product TTLa. If we make each xeL correspond to the element x = (xa)eL, L is mapped in a one-one linear way onto a linear subspace H of L. This mapping is a topological isomorphism, for if Up denotes the set of all y = (ya) in L for which yp = 0, the neighbourhood Ux of o in L is mapped onto the neighbourhood UanH of o in H, and the Ua form a sub-base of neighbourhoods of o for the topology of ULX. a 8. Dual spaces of subspaces and quotient spaces. If H is a linear subspace of L[3T[, every ueL defines a continuous linear functional u{0) on H, with the induced topology, and conversely the extension theorem shows that every continuous linear functional u{0) on H can be extended to a continuous linear functional defined on the whole of L. The linear mapping w->w(0) is called the natural homomorphism of L onto H'. The kernel of this mapping is H1, and so (1) The natural homomorphism of L onto H' defines an algebraic isomorphism L/H1 = H'. If u{0)eH' and ueL/H1 correspond to each other, u(0)y = uy for all yeH, so that (2) The induced topology £/s(L') and the topology £/s(H') coincide on H. The natural question, whether the natural homomorphism is a topological one, is answered for the linear weak topology as follows: (3) IfHczL is closed, the natural homomorphism of L onto H' is a topological homomorphism for the topologies £/s(L) and Hls(H). The algebraic isomorphism LIHL = H' is therefore a topological isomorphism for these topologies. By 7.(5) it is enough to show that the isomorphism L!/H1^H\ established in (1), is a topological isomorphism, when H' is given the topology %S(H\ and L/H1 the topology %S(L) induced by the canonical homomorphism K. The X/s(L)-neighbourhood F1 of L, where F is finite-dimensional in L, is sent by K into F1. The isomorphism (1) transforms %S(H) into a topology on L/H1, which has the sets G1 as base of neighbourhoods of o, where G runs through the finite-dimensional linear subspaces of H. F1 has the same image under K as V=F1 + H1. Since F1 is a neighbourhood of o, V is closed, by 2.(3), and is therefore orthogonally closed. Since V1 = F11 nH1JL = F nH, V is of the form G1. Thus we have shown that the image of every £/s(L)-
94 § 10. Linearly topologized spaces neighbourhood F1 is a 2/s(//)-neighbourhood G1; the topology Xls(L) on L/H1 therefore coincides with the topology Xls(H) induced by (1). If we apply (3) to H1 cz L' instead of to H, and use the fact that HL1 = H, we obtain (3 a) IfH is a closed linear subspace ofL, L/H = {HL)\ for the topologies Zls{L) and 2JH1). For the dual space {L/H)' of L/H, with the induced topology, we have (4) IfH is a closed linear subspace of L[£], {L/H)' is algebraically isomorphic to H1. Proof. If ueH1, a linear functional u' on L/H is clearly defined by u'x = ux. ux = 0 on a linear neighbourhood U of o, and so m'x = 0 on U = KU, where K is the cannonical homomorphism of L onto L/H. Thus u' is continuous on L/H. If, conversely, w'x is continuous on L/H, a linear functional is defined on L by ux = u'x. If u' vanishes on U, u vanishes on U = K{~1)U =) H, and so u is continuous and lies in H1. (5) The isomorphism {L/H)' = HL is a topological isomorphism for the topologies Xls{L/H) on {L/H)' and %S{L) on H1. For it follows from the fact that u'x = ux for ueH1, u'e{L/H)\ xeL and xeL/H that the £/s(L)-neighbourhood UXl,...,Xn in H1 corresponds to the £/s(L///)-neighbourhood Uxl,...,Xn in (L/H)', and conversely. (6) If L[£] is the topological direct sum LX@L2 of two closed subspaces, then L\ and L2 are Xls{L)-complementary to each other in L', and L\ = L'2 and L2=L\, for the linear weak topologies. Each xeL has a decomposition x = xx+x2, with xleLl9 x2eL2. If weL, a linear functional is defined on L by uxx = uxx. It is ^-continuous, for if ux = 0 for all xeU1@U2, where Ul and U2 are linear neighbourhoods of o in Lx and L2 respectively, then w1x = wx1=0, for all these x. Each u1 lies in L2, and clearly v = vt, if ueL^. Thus P2 = P holds for the linear mapping Pu = u1; it is a projection of L onto L2, and {I — P)L = L\. Pis 2/s(L)-continuous, forgiven x(1),...,x("} in L, where x^^x^ + x^0, xil)eLl9 x2i)eL2, it clearly follows from wxJ^O that w1x(I) = 0, for i=l,...,w. The decomposition L = LX®L2 is therefore a £/s(L)-topological direct sum, by 7.(6). The algebraic isomorphism L2 = (L/Ll)' follows from the isomorphism L/Lt = L29 for the induced topologies £, which was established in 7.(6). This is also a topological isomorphism for %S(L2) and Z^L/L^; applying (5), L2^L\ for the topologies %S(L2) and Xls{L).
9. Linearly compact spaces 95 The following example shows the difficulties associated with the idea of complementary spaces. Let Lx be the space <p®<p, consisting of all pairs (x,n) of finite vectors over K. Let L2 be the space of all vectors (u,o) = (x,n) + a(e,e), where (x,v))eLu c = {l,l,...} and aeK. (L2,L{) is a dual pair with respect to the bilinear form <(u,d), (x,t))> = ux + vi), where on the right hand side we take the scalar product of the vectors. Let H{ a Ll be the subspace consisting of all (x,o), and let H2 be the subspace consisting of all (o,n). (7) Ll=Hl®H2 is an algebraic complementary decomposition of L1 into two Zls (L2)-closed sub- spaces. The spaces orthogonal to H{ and H2 in L2 are H2 and H1 respectively, and so H\®H2=Lly which is a proper subspace of L2. By (6), the algebraic decomposition (7) is not %s (L2)-complementary; further the lattice-theoretic complementary decomposition L2 = H\y H2 in V(L2) dual to (7) is not an algebraic complementary decomposition. 9. Linearly compact spaces. Following Lefschetz, a linearly topo- logized space L is said to be linearly compact if every filter g with a base {Fa} of linear sub-manifolds Fa of L has an adherent point in L. Every compact linearly topologized space is clearly linearly compact, and so the concept "linearly compact" is weaker than "compact". However the most important properties of compact sets carry over to linearly compact subspaces. (1) A closed linear subspace F of a linearly compact space is again linearly compact. The proof of § 3, 2.(5) can be carried over to give (2) If A is a continuous linear mapping of a linearly compact space Lx into a linearly topologized space L2, then A{LX) is linearly compact. It follows from (1), (2) and 7.(3) that (3) // Lj is a closed linear subspace of the linearly compact space L, then L/Ll is linearly compact. Further, we have (4) A discrete linearly compact space L is finite-dimensional. Proof. Let L be infinite-dimensional, and let {xv}, vgN, be a basis of L. If Lv is the linear subspace spanned by the xv<, v' + v, Lv is closed, since L is discrete. The sets xv + Lv, vgN, and their finite intersections form a filter base of linear manifolds in L. An adherent point of this filter must be equal to Xl^v* which is not possible as there are infinitely many xv. v (5) The co-dimension of an open and closed linear subspace U of a linearly compact space L is always finite. Conversely, every closed linear subspace of finite co-dimension in a linearly topologized space is open.
96 § 10. Linearly topologized spaces If U is open and closed, U = 6 is open in L/U, and so L/U is discrete. L/U is linearly compact, by (3), and so, by (4), is finite-dimensional. Conversely, if U is closed and L/U is finite-dimensional, L/U is discrete, by 2. (5), and so U is open, as it is the inverse image of 6 under the canonical mapping. (6) A linearly compact space L is complete. A linearly compact sub- space of a linearly topologized space is therefore always closed. Proof. If <$={FP} is a Cauchy filter on L, then for each Ua there exists an Fm with x-yeUa, for all x, yeFm, and so Fm+Ua is a linear manifold xa+Ua. All these manifolds are closed, by 2.(3), and they form a filter base on L. By hypothesis there exists an adherent point jc0 in L; x0EFm+ Ua9 for each a, and so x0 + Ua = Fm+Ua => FPia\ x0 is therefore an adherent point of 5- The analogue of Tychonoff's theorem is valid: (7) The topological product of arbitrarily many linearly compact spaces is linearly compact. L is a linearly topologized space, by 2.(7). The proof of (7) is a transcription of the proof of Tychonoff's theorem given in § 3, 3.: Filters with bases of linear manifolds have a property similar to that for arbitrary filters: (8) Every filter with a base of linear manifolds can be refined to a maximal such filter. This is proved as in § 2, 7.(1). In place of §2,7.(3) we have the following result, proved in exactly the same way. (9) The image under a linear mapping of a filter maximal among those with a base of linear manifolds is again a filter maximal among those with a base of linear manifolds. (7) now follows from (8) and (9) in the same way as in Tychonoff's theorem. (10) Every continuous linear mapping A of a linearly compact space Lx[}t~\ into a linearly topologized space L2 is a topological homomor- phism with a closed image space. By (1), (2) and (6), the image of every closed linear subspace is closed. In particular A(LX) is closed, and by (3), L1/N\_A~\ is linearly compact. Under the one-one mapping A of L1/N\_A~\ onto A{LX) (which is continuous by 7.(4)), every closed linear subspace of finite co-dimension is sent into one with the same property, by (3). It follows from (5) that the X-image of every linear neighbourhood is a linear neighbourhood, so that A is open, and is a topological isomorphism. Thus, by 7.(5), A is a topological homomorphism.
11. The topology Z[k 97 10. £* as a linearly compact space. If the algebraic dual £* of a linear space E is given the topology Xls(E), £* is topologically isomorphic to Kd, with K discrete, by 6.(3). Since K, considered as a one- dimensional vector space, is linearly compact, 9.(7) shows that £* is linearly compact: (1) £* is linearly compact with respect to £/s(E). We shall establish a converse to this theorem, which will characterise linearly compact spaces. (2) // L[£] is linearly compact, X is the topology Xls(L). By 4.(4) it is enough to show that every ^-neighbourhood is a ^-neighbourhood. The linear ^-neighbourhoods of o have finite co- dimension, by 9.(5), and further are ^-closed subspaces, so that, by 4.(8), they are Xls(L) neighbourhoods. It follows from (2) and 8.(2) that (2 a) Every linearly compact subspace of a linearly topologized space L is linearly Xls(L)-compact. (3) // L[X~\ is linearly compact, Lis topologically isomorphic to (L')*, so that linearly compact spaces can be characterised as algebraic duals £* of linear spaces E, equipped with the topology 3^S(E). By 4, L can be considered as a subspace of L'*. The space orthogonal to L in L is [o], and so by 4.(6) the linear ^s(L')-closure of L in (L')* is equal to (L')*. But since L is complete, by 9.(6), L = (L)*. The rest of the assertion follows from (1). 11. The topology Hlk. Besides the weak topology, we can introduce another topology on L[£], and this we shall now investigate. n (1) The sum F= £ Ft of finitely many linearly compact subspaces of L is again linearly compact. n For by 9.(7) the topological product TT F{ is linearly compact; if / = i we map this by (x1,...,xn)-+xl +-** + xn onto F c L, we obtain a continuous linear mapping, and so, by 9.(2), F is linearly compact. (2) If U is a linear neighbourhood of o in L[£], U1 is linearly weakly compact in L. By 7.(7), L=U@LU where L{ is discrete, and so Lx=l*x. By 8.(6), L = U1@L\ and U1 is weakly isomorphic to L^; by 10.(1) and 8.(2), U1 is therefore linearly weakly compact. We now define the topology Zlk(L) in L[3f|, by taking as base of neighbourhoods of o in L the spaces C1 orthogonal to the linearly £/s(L)-compact subspaces C of L. By (1), the C1 form a filter base; 7 Kothe, Topological Vector Spaces I
98 § 10. Linearly topologized spaces the intersection condition of 2.(1) is satisfied, since Xlk(L) is finer than Xls{L), by (2). Thus L is a linearly topologized space under Xlk(L'). (3) Let L[X~\ be linearly compact. Then the topology Xlk(L) on L is equal to Xls(L), and Xlk(L) is the discrete topology on L. L and L are Uncomplete, and are dual to each other under Xlk. Proof. Xlk(L) is the discrete topology on L', as, by 10.(2), L is itself linearly weakly compact, so that L1 = [o] is a I/fc(L)-neighbour- hood in L. By 10.(3), (L[Xlk(L)])'= L. On the other hand, the method of proof of 9.(4) shows without difficulty that the linearly weakly compact subspaces of a discrete space are finite-dimensional; thus the topology Xlk(L) on L is equal to Xls(L). Completeness follows from 9.(6) and 10.(2), duality from 10.(3). If £ is a vector space and £* is its algebraic dual, the topologies Xlk on E and £* are therefore topologies which answer the question raised at the end of 6., and which maintain a complete symmetry between E and E*. The topology Xlk is characterised by (4) The topology Xlk(L) is the finest linear topology on L[X~\ which has L as dual space. Proof, a) Let *X* be a linear topology on L with the property that L[£*]'' = L[X"]'' = L. We must show that every linear ^-neighbourhood U is a ^-neighbourhood. U is £*-closed, by 2.(3), and so is orthogonally closed, by 4.(6). U is therefore the space orthogonal to U1, which is linearly compact, by (2); it follows that U is a ^-neighbourhood. b) In the opposite direction, we must show that every ^-continuous linear functional Uq is ^s-continuous on L. By hypothesis, there exists a linearly weakly compact subspace C of L', with the property that u0 vanishes on C1 c L. If we form the space C11 orthogonal to C1 in the algebraic dual £*, then u0 lies in C11. By 4.(6), C11 is the Xls(L)- closure of C in L*. On the other hand C is £/s(L)-complete, by 9.(6), so that CL1 = C, and u0eL. The topology Xlk is the analogue of the Mackey topology considered in § 21, 4., and (4) corresponds to the Mackey-Arens theorem (cf. §21, 4.). 12. Xlk-continuous linear mappings. We denote the space of continuous linear mappings from Z^IjXJ into L2[£2] by fi(L1p1],L2[22]). (1) Every continuous linear mapping A of L^X^ into L2[jX2] *s also Xls-continuous and Xlk-continuous, and we have (2) fiCMsa^M «=fi(L,[a;ta(L'1)],L2[a;ta(L'2-)]) = 2(L1[Zlk(Ll)lL2[Xlk(L2T\).
12. £Zfc-continuous linear mappings 99 The adjoint mappings A' form a linear subspace of £ (L'2 [%s(L2)l^ [Ita(L,)]) = £(L'2 [S^L,)],!/, [SJL,)]). Proof. As in § 9,4., it follows that a linear mapping is defined from L'2 into L'j by (3) (A'v)x = v(Ax) for all xeLl,veL2. This is the adjoint mapping A' of A, to which A is adjoint in turn. As in 5.(1), (2) follows from the weak continuity of A and A'. Further A is £//c-continuous, for given a linearly weakly compact subspace C of L'2, A'(C) is linearly weakly compact, by 9.(2), and the image Ax of each xeA'iQ1 lies in C1, by (3). The ^-continuity of each weakly continuous A' follows by interchanging the spaces and their duals. The fact, that we have equality instead of inclusion in (2) is proved by taking the topologies Xls (respectively Xlk) as the original topologies. It follows from (1) and 7.(6) that (4) Two closed linear X-complementary subspaces of L[X~\ are Xls(L)- and Xlk(L)-complementary. If F is a linearly compact subspace of L[I], F is linearly weakly compact, by 10.(2a), and F1 is a 2//c(L)-neighbourhood in L\ By 7.(7), this has a £//c(L)-complement H. By 11.(4) the Xlk(L)-dua\ space of L is L again, and by 8.(6) H1 is thus a closed space complementary to F in L. The projection P of Lonto F is ^-continuous: by 10.(2) a ^-neighbourhood in F is of the form U=UU, u nF, where the u, can be chosen in H. An element xeUUl Un, with decomposition x = xl+x2, xieF, x2etiL, satisfies the relations uix = uixl=0, so that PxeU. It therefore follows from 7.(6) that (5) Every linearly compact subspace F of a linearly topologized space L[X~\ has a X-complement. Information about the £/Jk-topology on L/H is given by the following result, corresponding to 8.(3): (6) // H is a closed subspace of L, the natural homomorphism of L onto H' is a topological homomorphism for the topologies Xlk(L) and Xlk(H), so that L/H1 = H' is a topological isomorphism. Similarly, L/H^iH1)' for the topologies Xlk{L) and Xlk(HL). The argument is analogous to the proof of 8.(3): a ^(^-neighbourhood of o in L has the form C1, where C is a linearly ^s(L')-compact subspace of L. A ^(/^-neighbourhood of o in L/H1 has the form D1, where D is a linearly £/s(i/')-compact subspace of H. By 8.(2)^D is a linearly £/s(L')-compact subspace of H, so that each D1 is a C1. Again, C1 has the same image as V = C1 + H1, and V is £//c-closed, 7*
100 § 10. Linearly topologized spaces by 2.(3), since C1 is a £/fc(L)-neighbourhood of o. Thus by 4.(6) V is orthogonally closed with respect to L. It follows from the fact that V' = CnH that V has the form D1, where D is linearly weakly compact in H. (7) // H is a closed linear subspace of L[3f|, the topology %lk(H') on H is finer than the topology Xlk(L). This follows by applying (1) to the inclusion mapping of H into E. The question whether %k(H') and Zlk(L) coincide on every closed subspace H of L[3f| depends upon whether every closed linear sub- space F of L for which L/F is linearly weakly compact has a linearly weakly compact complement, as does the validity of the result corresponding to 8.(5). We shall give a counterexample in § 31, number 4. 13. Continuous basis and continuous dimension. Using the idea of Cauchy filter, sums with more than countably many terms can be introduced in a linearly topologized space L[%~\. Let {xv}, veN, be n a family of elements of L. We form the partial sums £ xVi, for every » = i finite collection of different vh i= l,...,n. Let MV1 Vn denote the set n m of all partial sums £ xv.+ £ x^, as the v'k run through collections £=1 ' fc=l of pairwise distinct indices, different from the v,-. The sets MV1 Vn clearly form a filter-base on L. If this filter is a Cauchy filter, with limit x, x is called the topological sum of the xv, and we write x = £xv. If N is finite, we obtain the usual sum. v (1) // A is a continuous linear mapping from Lx into L2, and if x = Yjxv> tnen Ax = Y,Axv. V V Since this holds for finite sums, A(MVl ,...,VJ is the set MVi Vn of all n m Y, Axv.+ Y, Axv,k. Because of the uniform continuity of A (2.(9)), the £ = 1 ' k = 1 continuous linear image of a Cauchy filter-base is again a Cauchy filter-base, so that Ax is the limit of the filter defined by the MVl Vn. If follows directly from this that (2) // x = Yxv> tnen xa = £(xva), for each aeK. Further, if x = £xv and y = Y,yv> then tne family {(xv,j;v)}, vgN, is summable in LxL to (x,y); moreover the linear mapping (zl,z2)-+zl + z2 from LxL into L is continuous. It follows from this, using (1), that (3) // x = £xv and j> = ]£.yv in L, then x + y = £(xv + yv).
1. The duality of E and E* 101 A set {xv}, veN, is called a continuous basis of L if every xeL can be written in one and only one way as a topological sum x = £xv£v, £veK, and if the component £v is a continuous linear function of x, for each v. The smallest cardinal of a continuous basis of L is called the continuous dimension of L. It is invariant under topological isomorphism, by (1). If {xM}, //gM, is a basis for the vector space £, and if {u^} is the system dual to the xM in £*, {u^} clearly forms a continuous basis for £*, with respect to the weak topology. This is half of (4) Every weakly continuous basis of £* is the dual system of a basis of £, and conversely. If the family {u^} forms a continuous basis of E*, let xM be the weakly continuous linear functional on £* defined by uxfx = vfx, where u = Y,uiivn- un,xn = ® *f v'^V* and 1^x^=1. The xM are linearly in- k dependent, for given a linear combination £ x^.af (with not all the af vanishing), it is easy to find a £ u^.p^o, on which the linear com- i= i bination does not vanish. The linear span of the xM has o as orthogonal space in £*, and so it is equal to £ by § 9, 2.(1). The xM therefore form a basis of £, whose dual system is {u^}. If d is the dimension of £, every weakly continuous basis of £* therefore consists of d elements; i.e. the continuous dimension of £* with respect to Xls(E) is equal to the dimension of E. Using 8.(6), the continuous dimension of every closed subspace of £* is determined, as well. If we give the vector space E the discrete topology, every basis of £ is a continuous basis, and by the same token the dimension of a subspace of £ can also be considered as the continuous dimension with respect to the discrete topology. It is not known whether every linearly topologized space has a continuous basis. §11. The theory of equations in E and E* 1. The duality of E and £*. Let us gather together the most important of the results of the preceding paragraph. A vector space £ and its algebraic dual £* form a dual pair. If we give each of them the ^-topology, then this coincides with the discrete topology on £, and with the weak topology on £*. For these topologies, each space is the collection of continuous linear functionals on the other. If we make each closed linear subspace H of one of the two spaces
102 §11. The theory of equations in E and E* correspond to its orthogonal space H\ we obtain a dual isomorphism between the lattices V(E) and P(F*) of closed linear subspaces. Under this correspondence we have topological isomorphisms: If E = H1@H2, (1) (E/H.r^Hi^m, (2) • E/H^(Hiy^H2. If E* = G1®G2, (3) (£*/£!)' = G| = G'2, (4) E*/Gi =(<?!)* = G2. Proof. The first of the two isomorphisms in (1) and (3) (respectively in (2) and (4)) follows from § 10, 8.(5) (respectively 8. (3 a)), and the others follow from § 10, 8.(6). For this it must be borne in mind that if a space F has the discrete topology, then F' = F*. Remark. It is not necessary to assume that the direct sum E* = Gl@G2 of closed linear subspaces Gt and G2 is a topological sum: it follows from E* = Gl®G2 that E = G\®G\, by §9,3.(5). In the discrete space E algebraic and topological complements are the same, and so it follows from § 10, 8.(6) and § 10,12.(4) that the direct sum G|x © G2L = Gl@G2 is a Xls- and ^-topological direct sum. These relations, together with § 10,13, imply (5) If H cz E has dimension d and codimension c, then H1 has continuous dimension c and continuous codimension d, and conversely. There is also complete symmetry between the linear mappings Ae<5(E,F) and their adjoint mappings ^'eS(£*,F*). They form the collections £(£,F) and £(F*,£*) of all continuous linear mappings, for either %s or Xlk, from E into F and from F* into £* respectively. From §10, 9.(10) we have (6) Every weakly continuous linear mapping X'g£(F*,£*) is a topological homomorphism with a closed image space. By §9,2.(5), the same holds for Ae2(E,F). Since A'(F*) and iV[X'] have closed complements, by §9,3.(5), all the characteristic subspaces of A and A' are orthogonally closed. (7) The following relations hold between the characteristic subspaces of A and A': (8) Ar[A'] = A(£)1^(FM(£))*^C[>l]*, (9) A'{F*) = N[Ay^(E/N\_A])*^U[AY,
2. The theory of the solutions of column- and row-finite systems of equations 103 (10) n[a~] = A'(f*)1^(e*/A'(f*))'^C[A'J, (11) A(E) = N[A'Y^{F*/N\_A'~\)'^U[A'J9 and the isomorphisms are topological. Proof. It follows from v(Ax) = (A'v)x9 for xeE, veE*, that, for fixed v, A'v=o ifandonlyif veA(E)1, so that N\_A'~\=A(E)1. Fixings, we obtain that N\_A~] = A'(F*)1. Going over, in these equations, to the orthogonal spaces, and using (6), we obtain the first equations of (9) and (11) immediately. The isomorphisms of (8) to (11) follow by applying (1) and (3). If we call the continuous dimensions of A'{F*\ N|\4'] and E*/A'(F*) the rank r(A'), the nullity s(A') and the defect s'{A'), the next result follows directly from (9), (10) and (8). (12) The ranks, nullities and defects of A and A' are related as follows: (13) r(A) = r(A')9 s(A) = s'(A')9 s'(A) = s(A'). 2. The theory of the solutions of column- and row-finite systems of equations. Let A be a linear mapping from the vector space E of dimension d over K into the vector space F of dimension e over K. For given y{0)eF, the relation (1) Ax = y{0) is called a linear equation in xeE. Forgiven w(0)e£*, the equation (2) A'v = u{0) in veF* is called the transposed equation of (1); conversely, (1) is called the transposed equation of (2). (1) and (2) are called homogeneous if y{0) and u{0) are equal to o, respectively; otherwise they are inhomogeneous. The ideas which we have considered up to now allow us to describe the behaviour of the solutions of (1) and (2) in a way modelled on the determinant-free theory of the solution of finitely many linear equations in finitely many unknowns. By the rank, nullity and defect of (1) and (2) we mean the rank, nullity and defect of A and A' respectively; by considering 1.(8) to (12), we obtain the following theorem on the solution of equations: (3) The equations (1) and (2) have the same rank r, and the defect of either equation is equal to the nullity of its transposed equation. The solutions of the homogeneous form of equation (1) form a linear subspace of dimension s(A) = s'(A'\ and the solutions of the homogeneous
104 §11. The theory of equations in E and E* form of the transposed equation (2) form a closed linear subspace of F* with continuous dimension s'(A) = s(A'). The inhomogeneous form of equation (1) is solvable for just those y(0)eF which are orthogonal to all the solutions of the homogeneous transposed equation (2). The collection of all these y{0) forms a linear subspace of F of dimension r(A) and codimension s'(A). The inhomogeneous form of equation (2) is soluble for just those w(0)e£* which are orthogonal to all the solutions of the homogeneous equation (1). The collection of all these u{0) forms a closed linear subspace of F* of continuous dimension r(A) and continuous codimension s'(A'). All the solutions of the inhomogeneous equation can be obtained from just one solution, by adding all the solutions of the homogeneous equation. As a special case, if the homogeneous equation and the homogeneous transposed equation are insoluble, i. e. if they only have the trivial solution o, then the inhomogeneous equations have a unique solution for each y{0)GF or w(0)gE* on the right hand side. If we introduce a basis {xv}, veN, in £ and a basis {y^}, //eM, in F, then, by § 8, 5., (1) is transformed into e linear equations with d unknowns (4) Ia,v£v = C> A*eM, respectively 2a = n(0), V where 9I = ((a^v)) is a column-finite matrix and x = {£v}ecpd(K) and x)i0) = {rji^)}G(pe(K) are the vectors representing x and y(0) respectively. Using § 9, 4., the row-finite system of equations (5) ZaMv^ = y(v0)» veN> respectively 9Tt> = u(0), with o = {<p„}6coe(K), u<°> = {i/¥0>}etod(K), is obtained from (2) by considering the continuous bases of F* and F* dual to {xv} and {y^} respectively. (4) and (5) are again called systems of equations, each the transpose of the other. Conversely, any column- finite or row-finite system of equations can be interpreted as representing a linear equation (1) or (2) respectively, by means of a suitable basis, so that the theorem on the solution of equations can also be construed as a theorem about the solution of a row- or column-finite system of equations. 3. Formulae for solutions. The theorem on the solution of equations gives complete information about the behaviour of the solutions of column- and row-finite systems of equations over an arbitrary field K. Nevertheless, it does not enable us to represent the solutions of the
3. Formulae for solutions 105 given system of equations as functions of the matrix of the system of equations and the right-hand side. The theory of equivalence established in § 8 helps us to do this. If 21 is a column-finite matrix, then by § 8, 8. and 9. there exist in- vertible square column-finite matrices 23 and (£ for which (£2123 = X), where X) is a matrix for which the elements <V(v),v=l> while all the other terms vanish. Here v' runs through a index set l\T of r(2l) elements whose complement N" in the column-index set N contains s(2l) indices, while \i is a one-one mapping of l\T onto a set M' of indices, whose complement M" in the row-index set M contains s'i^X) terms. If we now introduce new variables in 2.(4) by putting x = 23 3, and multiply the resulting equation on the left by (£, we obtain (1) G^233 = D3 = ($;n(0) = t(0). The system (1) of equations is said to be equivalent to 2.(4); the solutions correspond to each other in a one-one way under the transformations £ = 233, 3 = 23_1x. Written out, (1) has the simple form (2) Cv' = tJ?U ° = tf}> v'eN'> A*'(v')eM', A*"eM". The homogeneous system has for solutions all those 3 which are of the form £ev»(V"> and which take arbitrary values on the v"eN", v" and these therefore comprise N[X)]. The inhomogeneous system only has a solution if all the 1^ = 0, for /i"eM". If this condition is satisfied, a solution is given by £<?> = t(m%, Cl—0, v'eN', v"eN". All the solutions are obtained from this by adding the solutions of the homogeneous system of equations. This result can be expressed by formulae. The transposed matrix X)' provides an inverse of X) in the sense of § 8, 4., for X) X)' is the matrix which has ones in the (fi'(v\fi'(v')) places of the main diagonal, and has zero everywhere else; X)X)' is therefore the projection onto the image- space of X). X)' X) is the projection onto the inverse-image space of X) consisting of all the 3 with Cv- = 0, for v"eN". (EN-X)'X) is the projection onto Af [X)], where (EN is the square unit-matrix defined on N. All the solutions of the homogeneous system (2) therefore have the form ((EN — X)' X))t), ne<pd(K), and a basis for the solutions is formed by the columns of (EN — X)' X). The inhomogeneous system is only soluble for those t(0) for which ((EM — X)X),)t(0) = o. One solution is then given by 3<°> = D't(0>. Going back to the equivalent system 2.(4), we obtain (3) // X) = (£$123 is the canonical form for 91, the non-zero columns of the matrix 23 ((EN — X)' X)) form a basis of the space of solutions of the homogeneous system SHx = o. The inhomogeneous system 9lat = t)(0) is
106 §11. The theory of equations in E and E* only soluble for those t)(0) which satisfy the condition ((£M — X) X)')Ctt)(0) = o. // this condition is satisfied, 3e(0) = 93 X)'(£t)(()) is a solution. By going over to transposed matrices, we obtain the following analogous result: (4) If I) = (£2123, 1)'= 93'2T (T is the canonical form for the row- finite matrix 9T. The non-zero columns of the matrix (£'(GM — T)T)r) form a continuous basis of all the solutions of the homogeneous system 91' o = o. The inhomogeneous system 9l't> = u(0) is only soluble for those u(0) w/uc/z sari5/> * fce condition (CBN - I)' T)) 93' u(0) = o. // this condition is satisfied, d(0) = (T X)93'u(0) is a solution. These formulae for the solutions are generalizations of Cramer's rule, although to be sure they provide no method for direct calculation, for it is not possible to determine 93 and (£ by an explicit calculation, since their existence is only established with the aid of Zorn's lemma. 4. The countable case. If E has countable dimension, the argument can be simplified considerably, and it is possible to give a constructive method which produces the canonical form in countably many steps. 00 Let E be equal to cp(K). A vector x= £ e,-^ is said to be of length n if {„ i=l is the last non-zero coordinate. If H1 is a linear subspace of cp(K), a subspace complementary to H{ is formed by the space H2 of all those i)e(p(K) for which rjk = 0 whenever k is the length of some vector xeHl. This easily established assertion (use a basis of H1 which contains, for each length appearing in H{, exactly one vector of this length) gives a construction for a complement in a way which does not depend upon Zorn's lemma. (1) In co(K) each weak accumulation point is the limit of a convergent sequence. For if u(0) is an accumulation point of the set M, there is for each n a u(f°eM which agrees with u(0) in the first n coordinates. Clearly u(n) converges weakly to u(0) [cf.§10,6.(4)]. Thus the use of filters is not necessary in co(K). If v„ is the first non-zero term in u = {u,-}eco(K), u is said to have shortness n. If Hj is a weakly closed linear subspace of co(K), let {Uj} be a set of vectors in H1 which contains, for each shortness; appearing in H^ just one vector Uj of this shortness. The {Uj} form a continuous basis of Hl9 and a closed complement of Hl is obtained by taking the set H2 of all continuous sums Xefc^fc> £fc£K, where k runs through the set complementary to the set of/s. k We now sketch a method of producing the canonical form of a row-finite matrix 2l = ((aik)), (i,k= 1,2,...,). A square matrix with an ordered index set is called a triangular matrix if the terms beneath the main diagonal vanish. The product of two triangular matrices is again a triangular matrix. If the columns c1,c2,..., of the matrix (£, considered as vectors in <p, have lengths 1,2,..., then (£ is an invertible triangular matrix, whose inverse is again a triangular matrix which can be calculated in a simple way. Let the first non-zero column of 31 have index s1 and length lu so that a/lSl=|=0. Let the next non-zero column be the s2-nd. We substract a multiple
5. An example 107 of the srst column from the s2-nd column, choosing the multiple so that the /rst element of the s2-nd column of the resulting matrix 2I(1) vanishes. If the s2-nd column of 2I(1) does not vanish, it has a length l2 different from lv For the second step, we subtract from the next non-zero column of 2I(1) (the s3-rd) suitable multiples of the srst and s2-nd columns, so that the /rst and /2-nd elements of the s3-rd column of the corresponding matrix 2I(2) vanish, and so on. The elementary operation which sends QI("-1) to 2I(n) is obtained by multiplying on the right by a matrix 93(n). This has ones in the main diagonal, and zeros every where else with the exception of the s„ + 1-th column, which has certain elements of K above the one of the main diagonal. The final result of these infinitely many operations is a matrix $, in which all the non-zero columns a7 have different lengths /,. We can adjoin certain cfc to the non-zero d7, to obtain a system of vectors in which each length appears just once. The matrix obtained in this way can be turned into an invertible triangular matrix by permuting its columns, and is therefore itself invertible. We denote its inverse by (L This sends each d7- into an cz., so that the matrix (£$[ has at most one one in each column and row. By permuting the rows or columns if necessary (again by multiplication by invertible column-finite matrices) this can be transformed into a canonical form, which is a unit matrix when the zero rows and columns are deleted. To complete the proof of our assertion, it remains to show that 31 = 9193, where 93 is an invertible triangular matrix. It is natural to take 93 as limit of the sequence 93(1),93(1)93(2),.... A sequence $R(n) of matrices in 2(<p) is said to converge to the matrix We£((/>), if to each k there corresponds an n0(k), such that the first k columns of W — Wn) vanish, for n>n0. To each Cauchy sequence $R(n), defined in a corresponding way, there clearly exists a column-finite W as limit. For this concept of limit, S = lim 2I(n), and the sequence of products 93(* ]... 93(n) converges to an invertible triangular matrix 93. The missing relationship S = limE(n) = limE©(1\..©(n) = E93. follows from the easily established rule for column-finite matrices, that if lim SR(n)=W and limSW^aR, then lim«(n)aR(n) = 9?aR. The multiplication of # by £ can also be obtained by making infinitely many elementary operations on the rows of 21, and these operations can be made alternately with the column operations. It finally remains to confirm that this infinite product of elementary operations converges in the sense described above. 5. An example. Show that the matrix / 1 2 2 2 21 = f 1 1 1 1 0 0 0 1 0 0 0 1 0 0 0 0 i 0 0 0 0 2 2 2 1 0 0 2 2 2 1 0 0 2 2 2 1 1 1 2 2 2 2 2 1 :\ is equivalent to the matrix which is obtained from the unit matrix (E by inserting zero-columns between the second and third columns, the fourth and fifth, and so on. The matrices 93 and (£ which make the transformation can be determined easily, and so formulae can be obtained for solving the system of equations with matrix 21.
108 § 12. Locally linearly compact spaces § 12. Locally linearly compact spaces 1. The structure of locally linearly compact spaces. We now consider a further example related to the theory of linearly topologized spaces developed in § 10. A linearly topologized space L[X] over K is said to be locally linearly compact if it has a linearly compact neighbourhood V of o. By § 10, 7.(7), there is a discrete ^-complement W of V. The topology X on W is the discrete topology, and so also is the topology induced by Xlk(V'\ since Xlk{V) is finer than X. By § 10,10.(2) and 11.(3), X is the same as the topology Xlk(V) on V, and this coincides with the topology induced by Xlk(L\ since on the one hand this last topology is finer than that induced by X, and on the other Xlk(V) is finer than Xlk(L) on V, by § 10,12.(7). By § 10,12.(4), V and W are also ^(^-complementary, and so both Xlk(L) and X are direct sums of the topologies Xlk{L) onFandPy,sothat X = Xlk(L) on L. Using § 10,10.(3), § 10, 6.(5) and § 5,7., we obtain (1) Every locally linearly compact space L[X] is topologically isomorphic to a topological direct sum (pdl(K)®cod2(K). The topology of L is the topology Xlk, and it is equal to the product of the discrete topology on <pdl{K) and the weak topology on cod2(K). Every locally linearly compact space is complete. Conversely, we clearly have (2) The topological direct sum L = Ll®L2 of a linearly compact space Lx and a discrete space L2 is locally linearly compact. Lx is an open and closed neighbourhood of 0 in L. A basis of L2 and a continuous basis of Lx together comprise a continuous basis of L1®L2. The dimensions dx and d2 are not in general uniquely determined, for finite-dimensional subspaces of <pdl can be taken across to codv and conversely. If dt or d2 is finite, we can therefore always take a decom- sition in which one of the summands vanishes. Nevertheless, we have (3) Suppose that L is locally linearly compact, but neither discrete nor linearly compact. If L = Ll® L2 is a decomposition into a discrete space L1=cpdl and a linearly compact space L2=a>dl, the infinite dimensions dt and d2 are uniquely determined. For if L — Ll®L2 is another such decomposition, with L1=(pd>l and L2=a>d>2, then (L2 + L2)/L2 is linearly compact in L/L2 = Ll, by §10,9.(2), and is therefore finite-dimensional by §10,9.(4), It follows that d'2^d2, and that d'2 = d2. The image of 1^ under the canonical
2. The endomorphisms of \j/ 109 mapping of L = Ll®L2 onto L/L2 = Ll contains a complementary space of (L2©L2)/L2; thus d\^dl, and so d\=dl. o)dl@cpd2 is the dual of cpdl®a)d2, and so we obtain the following theorem of Lefschetz: (4) The dual space L of a locally linearly compact space L is again locally linearly compact with respect to the topology Xlk9 and (L)' = L. If dl=d2, L is ^-isomorphic to L. (5) Every closed linear subspace F of a locally linearly compact space L[%~\ is locally linearly compact, and has a %-complementary space G. Proof. Let L = LX®L2, with Lx discrete and L2 linearly compact. F2 = FnL2 is a linearly compact neighbourhood of o in F in the induced topology, so that F is locally linearly compact. By §10,7.(7), F = Fl®F2, with Fj discrete. F2 has a topological complement G2 in L2, by §10,12.(5), and by §7,6.(2) F + L2 = F1®F2@G2. Hence by §10,7.(7) the sum FX®L2 is a topological direct sum. Again by §10,7.(7), F + L2 has a discrete topological complement Gu and L = Gl®(F1®(F2®G2)). Thus L is the topological product of these four subspaces. As the order and way in which they are put together does not matter, L = (Fl® F2)®(G1®G2) = F®G is the required topological complementary decomposition. 2. The endomorphisms of *//. In the case where both dimensions dl and d2 are countable, we obtain the semifinite space \j/ = (p®aj. Its dual space iJ/' = co®(p is topologically isomorphic to ijj. A full account of the equivalence of the endomorphisms of if/ can also be established. Using this, the theory of the corresponding "semi-finite" systems of equations is at our command, in the same way as for row- and column-finite systems of equations. First we show that every continuous mapping Ae2(ij/) can be taken in canonical form, by choosing two continuous bases suitably. By 1.(5), the kernel N\_A] has a ^-complementary inverse-image space U\_A] which is again locally linearly compact. By §10,7.(6), ijj/N\_A] is topologically isomorphic to C/[X], and, by § 10, 7.(4), U\_A] is mapped continuously by A in a one-one way onto G = A(ip). Let G be the closure of G. By 1.(5), G has a ^-complement Gx. We decompose U\_A~\ as Fl®F2, where F\ is discrete and F2 is open and linearly compact; F2 can be taken either as o or isomorphic to co. The image space y4(F2) = G2 is again linearly compact, by § 10, 9.(2), and by §10,9.(10). A establishes a topological isomorphism_between F2 and G2. By 1.(5), G2 has a topological complement G2 in G. The inverse images of those elements of A(ijj) which lie in G2 form a linear
110 § 12. Locally linearly compact spaces subspace F2 of U\_A]. By §10,7.(7), F2 is topologically complementary to F2 in l/[4], and is discrete, so that U\_A] = F2®F2. Since G = A(F2)®A(F2) is dense in G = G2®G2, and since A{F2)^G2 and A(F2) = G2, A(F2) is dense in G2, by §1,8.(1). We now consider the continuous one-one mapping from the discrete space F2 into G2. Let G4 be open and linearly compact in the locally linearly compact space G2. Once again, we can take G4 either to be o or to be isomorphic to co. By § 7, 6.(5) and § 10, 7.(7), there is a discrete topological complement G3 to G4 in G2, and so A(F2) = (A(F2) n G3) © (A(F2) n G4) = H3 0 H4. Let F3 and F4 be the inverse images of H3 and H4, so that F2 = F3®FA, H3 = A{F3\ HA = A(F4). Since A(F2) is dense in G2, we must again have A{F3) = G3, A(FA) = GA. G3 is discrete, so that A determines an isomorphism of F3 onto G3. It finally remains to investigate the linear mapping A from the discrete space F4 into the linearly compact space G4, in the case where G44=o. Equivalently, we may consider a continuous one-one linear mapping A from <p into co, whose image space is dense in co. We now refer to the ideas developed in §11,4. Let x1,x2)... be a basis of cp. Since A(cp) is dense in co, there is, for each /c= 1,2,..., a zecp whose image y = Az has shortness /c. In particular let zl be chosen so that yi=Azt has shortness 1. If xt is linearly independent of zy (otherwise, take x2), there is a zkx=xl-\-zlXll in cp whose image ykl has shortness fe^l. If fci>2, there exist elements z2,...,zfcl_1 in ^ whose images j;2, ...,ykl_i have shortness 2,...,kl — 1 respectively. The elements z j,..., zfc are linearly independent, since their images clearly are. Let xi2 be the first term in xl5 x2,... which is linearly independent of z1,...,zkl. Then there exists a ^2 = xi2 + z1>l21 + -" + zfci>l2fci, whose image ykl has shortness /c2>/c1. If/c2>/c1 + l, elements zfcl + 1,..., zfc2_! of cp can again be inserted so that all the elements z1?..., zkl are linearly independent, and so that their images have shortness 1,..., /c2, respectively. Let xl3 be the first of the terms xi2 + l9... which is linearly independent of zl5..., zk2. By continuing this process we obtain a basis {zj of cp for which the images y{ = Az{ have all possible different shortnesses, so that the {jj form a continuous basis of co. Combining everything together, corresponding to AeQ(iJ/) we have constructed two direct topological decompositions of ip (1) ip = N[A]®F2®F3®F^ (2) il/ = Gt ®G2®G3®G4 with the following properties: the null-space N\_A~\ can be of finite dimension n or can be topologically isomorphic to cp,co or ij/. The same holds for the complement Gx of the closure of the image space A(\jj).
3. The theory of equivalence in \J/ 111 F2 and G2 are either both o or both isomorphic to co, and A defines a topological isomorphism of F2 onto G2 which takes the form Ax2l) = y2l\ i=l,2,..., when continuous bases x2l) and y2l) of F2 and G2 are chosen in a suitable way. F3 and G3 are both discrete, and have the same dimension, which can be finite or countably infinite. A defines an isomorphism which takes the form A x^ = y{^\ when bases x(30 and y^ of F3 and G3 are chosen in a suitable way. Either F4 and G4 both vanish, or they are isomorphic to cp and co respectively. By choosing a basis x4° of F4 and a continuous basis y4l) of G4 suitably, X is represented as the mapping Xx4° = y4° from F4 onto a dense subspace of G4. If U\_A~] is infinite-dimensional and F3 is finite-dimensional, F3 and G3 can always be taken to be o. If A is a mapping of infinite rank, it follows from this account that the mapping from U\_A\ onto A(ijj) is of one of the seven following types (we denote by q> the space q> considered as a dense subspace of co, and provided with the topology induced by the weak topology on co; this in distinction to q>, with the discrete topology): (3) U[A~\: cp, <p, <p(B<p, co, 0)®<p, w®cp, w®cp®cp, A(ip): <p, q>, (p@(f>, co, co©cp, co©<p, w®cp®q). 3. The theory of equivalence in ij/. A continuous basis of \jj is obtained if \j/ is decomposed in some way as a direct sum \j/l ® \\j2, with ij/l discrete and \j/2 open and linearly compact, and if an algebraic basis x_l5 i = l,2,..., is chosen for ^x and a continuous basis xh / = 1,2,..., is chosen for \j/2. If a second continuous basis x'_h x\ is chosen, the mapping Bxk = x'k, (fe= ±1, ±2,...) is a topological isomorphism of ip. Two mappings Al,A2e2(\J/) are said to be equivalent if there are topological automorphisms B and C for which A2 = BAlC; the next two results then follow without difficulty from the results of the preceding number and the remark made above: (1) Two mappings of finite rankCi.e. N[X] has finite codimension in \j/J are equivalent if they are of equal rank. (2) Two mappings of infinite rank are equivalent if a) their null-spaces and the complements of the closures of their spaces are respectively either of the same finite dimension or of the same type <p, co or ij/, and b) the dimensions of the spaces F2, F3 and F4 are in each case either both 0 or both oo. This is a complete division into equivalence classes, as the next result shows:
112 § 12. Locally linearly compact spaces (3) Two mappings are equivalent if and only if the conditions given in{\) and (2) are satisfied. This is immediately obvious for (1). In case (2), which is clearly not equivalent to (1), we argue as follows: BAX C has null-space C~ ^(N^A^]), which is of the same type as Nf/lJ, and in the same way B sends a complement G[l) of A^ip) into a complement BG[l) of BA^^), so that these are both of the same type. The inverse-image space contains an infinite-dimensional linearly compact space if F2 is infinite-dimensional, and is discrete if F2 is o; these properties are preserved under equivalence. If F3 is infinite-dimensional, the mapping sends an infinite-dimensional discrete subspace onto one of the same kind, and if F3 is zero, this is not the case. If F4 is infinite-dimensional, A(\j/) is not closed, whereas A(ip) is closed if F4 is o. By using similar ideas, it is easy to see that of the seven types 2.(3) of image spaces A(\J/) no two are topologically isomorphic; we can therefore express this result about equivalence in the following terms: (4) Two continuous linear mappings of ip into itself are equivalent if and only if their null-spaces, their image spaces, and the complements of the closures of their image spaces are topologically isomorphic. If we have an underlying continuous basis ...,x_2,x_1,x1,x2,... of ij/, a linear mapping Ae2(ij/) is represented by a matrix 2l = (alfc)), which is infinite in all directions, and which has the following form: • a_2_2 • a_!_2 . ax_2 . a2_2 a-2-1 a-i-i «i-i a2_i a-21 a-n an a2i a_22 . a-12 • a12 . a22 «, «3 <U2 «4 where 2^ is column-finite, 2l2 is finite (and in particular is column- and row-finite), 2l3 is arbitrary and 2l4 is row-finite. Such a matrix is said to be semi-finite. The fact that 2l2 is finite follows easily from the condition that A maps \jj into itself. In addition to the topological automorphisms of \J/, which are all equivalent to the identity mapping, there is just one equivalence class of mappings A for which both Ax = 0, xeip, and Afu = 0,ueiJ/\ are not soluble. This is the case where N\_A~]=o, Gl=o, and Fl9 F2 and F3 are all infinite-dimensional, i.e. the last case of 2.(3). By numbering the ele-
1. Linearly bounded subspaces 113 ments of the bases in a suitable way, we obtain the following normal form: 1000000 000 0010000 000 0000100 000 0000001 000 0000000 100 0000010 000 0000000 010 0001000 000 0000000 001 0100000 000 We leave it to the reader to transform the theory of equivalence into a theory about infinite equations represented by semi-finite matrices, and with semi-finite vectors as solutions (cf. Kothe and Toeplitz [1]). J. Dieudonne [4] has shown that there is no corresponding simple theory in the uncountable case. § 13. The linear strong topology 1. Linearly bounded subspaces. We continue with the general theory of linearly topologized spaces, by introducing an analogue to the concept of bounded set; this concept is particularly important in the later theory. Let L be a vector space over K with a linear topology X, which is defined by linear neighbourhoods Ua, aeA, of o. A linear subspace F of L is said to be linearly ^-bounded if (F + Ua)/Ua is finite-dimensional for each Ua. If X is the discrete topology, the linearly ^-bounded subspaces are precisely the finite-dimensional subspaces; if X is the linear weak topology Xls(L), every linear subspace is linearly weakly bounded, since each Ua is of finite codimension in L. (1) The closure of a linearly X-bounded subspace F is again linearly bounded. By §10,2.(3), F+L/a is closed, so that F11-^ t/a = F + Ua follows from F+U^iF+U,)11 => F^+U^ F+Ua. Thus if (F+ UJ/U, is finite-dimensional, so is (F±L-{- C/a)/C/a. 8 Kothe, Topological Vector Spaces I
114 § 13. The linear strong topology (2) The sum of finitely many linearly X-bounded subspaces of L is again linearly X-bounded. Every linear subspace of a linearly X-bounded subspace is again linearly X-bounded. The proof is trivial. (3) The continuous linear image in L2[£2] °f every linearly X-bounded subspace of L^X^ is linearly X2-bounded. Let Ae£(Ll9 L2). If V is a closed linear neighbourhood of o in L2, there is a neighbourhood U of o in Lx with A(U)a V. Since (F+U)/U is finite-dimensional, A(F+U)/A(U) = (A(F) + A(U))/A(U) is finite-dimensional, and so therefore is (A(F)-{- V)/V. (4) A linear subspace F of L[3f| is linearly X-bounded if and only if X and the linear weak topology coincide on F. a) If X = Xls on F, then for each ^-neighbourhood Ua there exists a ^-neighbourhood Kwith Fn V=Fn Ua. Then by § 7, 6.(6), (F + t/a)/t/a ^F/(FnUJ = F/(FnV)^(F+V)/V. Since L/V is finite-dimensional, so therefore is (F+t/a)/t/a. b) Conversely, if (F+Ua)/Ua is finite-dimensional, we can write F+l/a=l/a©[x1,...,xII], where the x,- are linearly independent. Because each L/a© [xl5..., x,_i, xf+1, ...,xn] is closed, there exist continuous linear functionals ul9...,un in L which vanish on C/a, and which take the values M.xf=l,Mf(xk)=o for i#=fe, on the xh by § 10, 4.(1'). We than have F+ K= V® [xl9..., xn] for the weak neighbourhood K=l/M1,...,Mn(o), so that (F+K)/K^(F+(7a)/(7a, and F/(FnK) = F/(FnUJ; since F/FnV is finite-dimensional and Fnt/ac=FnK, we finally deduce that F nU0L = F nV. Starting from a dual pair {L2,Ll}, linear £/fe-boundedness is defined in both Lx and L2. This is the most important of the various concepts of boundedness. (5) Every linearly weakly compact subspace F of L[X] is linearly Xlk-bounded. If U is a linear ^-neighbourhood of o in L, L/U is discrete, by § 10, 7.(2). Since (F+ U)/U is the continuous image of F, and is therefore linearly weakly compact, it must be finite-dimensional. The converse of (5) is not valid. Consider the dual pair (q>,q>y. By § 10, 10.(3) and §9,5.(3), a linearly weakly compact space never has countable dimension. Thus only the finite dimensional linear subspaces of q> are linearly weakly compact. Hence Xlk = Xls, and q> itself is linearly £/fc-bounded. 2. The linear strong topology. Once again, let ^2^^ be a dual pair. As B runs through all the linearly £/fe-bounded subspaces of L2, the spaces B1 satisfy the conditions of §10,2.(1), by 1.(2); taken as
3. The completion 115 neighbourhoods of o they therefore define a linear topology on Ll5 which 'we call the linear strong topology %lh(L2) on L1. It follows immediately from 1.(5) that (1) The strong topology %h(L2) is finer than the topology %k(L2). Further, it follows from § 10,11.(4) and 1.(1) that (2) The dual space of Lx [%lh(L2)~] is equal to L2 if and only if %lh(L2) = %lk(L2\ and so if and only if every weakly closed linearly Unbounded subspace of L2 is linearly weakly compact. Let L be a linearly topologized space with topology X. The dual space L, equipped with the strong topology Zlh(L)9 is called the strong dual ofL. Lis said to be strongly semi-reflexive ifthedual [L'ljX^])' of the strong dual is equal to L; Lis said to be strongly reflexive if, in addition, the strong topology Zlb(L) on L is the same as the original topology. (3) L is strongly semi-reflexive if and only if every closed linearly Unbounded subspace of Lis linearly weakly compact. (4) L is strongly reflexive if and only if every closed linearly Xlk- bounded subspace is linearly weakly compact, in both L and L, and the original topology H is the same as %lk(L). Proof. (3) is a consequence of (2). If L is strongly reflexive, then % = %lh(U), and a fortiori <Xlk(L) = <Xlh(L), so that every closed linearly £,fc-bounded subspace is linearly weakly compact in L as well. The converse is also clear. As an application, we give a refinement of the theorem of Lefschetz (cf. § 12,1.(4)): (5) Every locally linearly compact space L is strongly reflexive. By § 12, 1.(1) and § 12, 1.(4), it is enough to show that every linearly £/fc-bounded subspace B of L is contained in a linearly weakly compact subspace. We can write L=LX © L2, with L{ discrete and L2 open and linearly weakly compact. By 1.(3), (B + L2)/L2 is linearly £,fc-bounded in the discrete space L/L2. Thus (B + L2)/L2 is finite-dimensional, so that B + L2 is linearly weakly compact. 3. The completion. In § 5, 5, it was shown how a uniquely determined completion can be constructed for every Hausdorff uniform space. Later (§ 15, 3.), we shall see that the completion of a topological vector space constructed in this way is again a topological vector space. In the present case, however, there is a simple way of constructing this completion, which corresponds to the method of Grothendieck (cf. § 21, 9.). This construction is independent of the ideas of § 5. s^
116 § 13. The linear strong topology Let L be a vector space with topology X defined by weakly closed linear neighbourhoods U ofo. We denote by Lthe collection of all those linear functionals y on L whose restrictions to each U1 are weakly continuous. L is clearly a vector space which contains Las a linear subspace. We define a linear topology X on L by taking as neighbourhoods in L the linear subspaces U11 => U orthogonal to the spaces U1. It is easy to see that X is a linear topology: the intersection of finitely many JJtL contains the neighbourhood (f]^i) , so that the U11 form a filter base. That f] U11 = o follows from the fact that a yeL which vanishes on all of the U1 is identically zero. Since U11 n L= [/, X is the topology induced on L by X. That L is the completion of L now follows from (1) Lis complete with respect to X, and L is dense in L. If {Fa} is a Cauchy filter on L, then for each U1 there is an Fa such that all the yeF^ have the same restriction yffl to U1, and this restriction is a weakly continuous linear functional on U1. The y$ defined on all the sets U^ in this way determine a linear functional y0)eL, defined on the whole of L. For if U\ and U% are distinct, and if yi=y{ift for all y^eF*1 and y2=y{Sl for all y2eF*2 respectively, then we clearly have y^=y = y^) on U\nU% for all yeFyczFainFa2. The linear functionals yffi and y(^i therefore agree on the intersection of their domains of definition, and y{0) is the limit of the filter {Fa}. Since ytfl, being a weakly continuous linear functional on U1, can also be defined by an veL, j;(0) is a 5^-adherent point of the x^, and so L is dense in L. As a first consequence, we obtain a further characterization of the linearly ^-bounded sets. A linear subspace F of L is said to be linearly 2-precompact if the closure of F in L is linearly X-compact. (2) A linear subspace F is linearly X-precompact if and only if it is linearly X-bounded. It is sufficient to show this for L itself, for a linear subspace F of L[X] is linearly ^-bounded or linearly 2-precompact if and only if it has the same property with respect to the topology induced on F by X. If Lis linearly ^-bounded, L/U is finite-dimensional, and so by § 10, 8.(4) U1 is finite-dimensional; the linear functionals in L therefore consist of all the linear functionals on L', i.e. L=(U)*. L is therefore linearly weakly compact, by § 10,10.(1), and X = Xls{L).
4. Topological sums and products 117 If, conversely, L is linearly 2-precompact, L is linearly ^-compact, and X is the weak topology, by § 10, 10.(2). If therefore follows from the remark before 1.(1) that L is linearly X-bounded. If L is :X/fc-complete, every closed linearly :X/fc-bounded subspace of L is linearly ^-compact, and is therefore linearly weakly compact, so that by 2.(3) we have (3) // L is Xlk-complete, then L is strongly semi-reflexive. Likewise, we have (4) // L and L are X^-complete, then L[^Xjfc(L')] is strongly reflexive. Both these theorems are no longer valid if we only require X//rcompleteness, as the example of the dual pair <</>,</>> shows (cf. the remark at the end of 1.). In both spaces the strong topology is the discrete topology, under which each is complete, while the strong dual spaces are isomorphic to co. Let us mention one closely related problem: is a linearly topologized space strongly complete if every closed linearly X/k-bounded subspace is linearly weakly compact? 4. Topological sums and products. We give some examples to illustrate the theory we have just developed. Let La, aeA, be linearly topologized spaces with topologies Za. (1) If L= ® La is the topological direct sum of the spaces La, then L a is Xlk-isomorphic to TT L^[2zk(La)]. a // L = TT La is the topological product of the spaces La, L is Zlk- isomorphic to © L'a[Zlk(La)]. a Proof. Let us denote by % the topology defined on L=® La by the topologies Za, as in § 10, 2. We can consider La as a linear subspace of L. Each ueL then defines a ^-continuous linear functional uaeL'a on La, and this we can consider as an element of L', by making it take the value zero on the spaces Lp with /?=j=a. It is now quite clear that the correspondence w^{wa}eTTL/a is an algebraic isomorphism between the spaces L and IT L'a. a In a similar way, for the second half of the proposition we establish the algebraic isomorphism L = © L'a (for a neighbourhood of o in a TTLa has the form IT Ua, with Ua = La for all but finitely many a, so a a that if ueL there are only finitely many wa=No). We show that the isomorphisms are ^-isomorphisms in the following way: the projection Ca of a linearly :X/s(L')-compact set C of © La or of IT La onto La a a is linearly :X/s(L^)-compact, by § 10, 8.(2). In the case L = TT La, on the a one hand C is contained in the topological product of the linearly
118 § 13. The linear strong topology £js(Z4)-compact sets Q, and on the other hand TT Ca is again linearly :X/s(L')-compact, by §10,9.(7), so that the topology Xlk(L) on L is equal to the direct sum topology of the spaces La\%lk(L„f\. In the case where L= © La, the homeomorphism is established in a a similar way, using the fact that only finitely many projections Ca are different from o, so that C is contained in © Ca. For this, it is sufficient i ' oo to consider the following case: Let L= © Lk, and suppose there is oc fc=i a sequence x{i) = £ xjt°, x[i]eLk, in C, with xj-04=o, and x^ = o for k= i /c>i. The linear span H of all the x{l) is infinite-dimensional. We show that H, considered as a subspace of L, is discrete. There is a neighbourhood Ul ofo in Lt for which Lx = Ul © [x(1)]. The coset in (Lt © L2)/Ll = L2 determined by x{2) is non-zero, and so there is a neighbourhood U2 of o in L2 for which {L1 © U2)n [x(2)]=o. Thus Lj ©L2= [/j © U2 00 ©[x(1),x(2)], and so on. Then the neighbourhood © Uk of o in L meets H only in the point o. k=1 H, being a closed linear subspace of C, is linearly compact, and is therefore finite-dimensional, by § 10, 9.(4); this gives a contradiction. We know (§ 5, 7.) that IT La is ^-complete if the spaces La are com- plete with respect to £a. Similarly we have (2) The topological direct sum L = © La of complete spaces La is complete. Proof. If <&={F{P)} is a Cauchy filter on L, the projection Pag={Pa(F(/J))} is a Cauchy filter on La. Since La is complete, Pag has a limit xaeLa. We assert that xa = o, except for at most finitely many a,. Otherwise there would be a sequence xa.4=o, i=l,2,.... Let (7ai be a closed linear neighbourhood ofo in La., with the property that o does not lie in x +U . Let U = © [/a be a neighbourhood of o in L for a which the Ua. are the neighbourhoods which have just been determined, and for which the other Ua are chosen arbitrarily. Further, let F{(i) be small of order U. If now zeF{P\ Poc.z = z0ii lies in xa.+ Ua., and so must be different from zero for all i, which is impossible. Since Fip) cannot be empty, we have reached a contradiction. If xai,..., xan are the finitely many non-zero limits of the filters Pa5, then clearly xai + ••• + xan is the limit of g. (3) // the spaces La ar^ all strongly reflexive, then © La and TT La ar^ strongly reflexive. Proof. La is strongly reflexive, and so the topology of La is the 2/fc-topology. Applying (1) to L = © L'a[Zlk(Lj] instead of to L= © La,
5. Spaces of countable degree 119 we obtain, for L = (L)\ that 17 La has the topology %k(L!). In the a same way we show that the direct sum topology on © La is equal a to Xlk. Our assertion therefore follows from 2.(4), provided that we can show that every closed linearly ^-bounded subspace of 17 La a (respectively © La) is linearly weakly compact, when each La has this property. Let C be closed and linearly £/fc-bounded in 17 La. (C+U)/U is finite-dimensional, for each ^-neighbourhood U = TlUa in TTLa. If Ca is the projection on La, it follows that (Ca+Ua)/Ua is finite-dimensional, and by 1.(1) the £/fc-closure Ca is also linearly bounded. By hypothesis, Ca is linearly weakly compact, by § 10, 9.(7) 17 Ca is, as well, _ a and finally so is Cc TTCa. If C is closed and linearly ^-bounded in a ©La, only finitely many Ca can be non-zero, for otherwise a neigh- bourhood U = © Ua could be constructed, as in the proof of (1), for a which (C+ U)/U would be infinite-dimensional. The linear weak compactness of C then follows as before. 5. Spaces of countable degree. Starting from the spaces cp, co and \jj we construct further spaces with countable continuous bases, which are no longer locally linearly compact, but which are strongly reflexive. By 2.(5), cp and co are strongly reflexive. We denote them by Sx and Si respectively. Let S2 be the topological direct sum of countably many copies of Si. Clearly all the vectors of S2 are obtained when in each vector {^l9..., f„,0,0,...} of cp the non-zero cjf are replaced by arbitrary nonzero vectors u, of co, and the zeros are replaced by the zero-vector o of co. Because of this, we also denote S2 by cp co. By 4., S2 is the topological product of countably many copies of S'[ = cp. We also write cocp for S2, since the vectors in S2 can be obtained as above by replacing the coordinates of vectors in co by vectors in cp. This procedure can be extended. Let a run through the ordinals of the first and second classes. We suppose that Sp has already been defined, for /?<a. If a is not a limit ordinal, we put Sa equal to the topological direct sum of countably many copies of S'^j, so that Sa = cpS,a_1. If a is a limit ordinal, we put Sa = © S«. P<0L By transfinite induction, we easily obtain (1) Sa is %lk-complete and strongly reflexive. S^ is equal to coSa_1? if a is not a limit ordinal, and S'a is equal to TT S«, if a is a limit ordinal. P<oc
120 § 13. The linear strong topology We say that a linearly topologized space over K is of countable degree if it is obtained from cp(K) and co(K)by a well-ordered collection (up to ordinals of the second class) of topological sums and products. It can be shown (cf. Kothe [1]) that every space of countable degree can be transformed into one of the spaces Sa, S"a or Sa@S'a by a permutation of countably many coordinates of its vectors, and that these canonical forms are all different, in the sense that no two are topological^ isomorphic. Thus the degree a of such a space can be defined; this is an invariant of the space. Only the spaces cp, co and \jj of the first degree are locally linearly compact. In all these spaces, the unit vectors form a continuous basis. In (i) co there is no simple theory of equivalence similar to that for cp, co and ij/9 and no full account has yet been given. 6. A counterexample. Circumstances can be considerably more complicated for general linearly topologized spaces than they are for cp and co. This is made particularly clear by the example of the strongly reflexive ^-complete space cpco©cocp of second degree (cf. Kothe [5]). We assume in the following that K does not have characteristic 2. By 5., the vectors in cpco and cocp have the form We denote the vector with £ik= 1, and ^ = 0 otherwise, by el7c. A vector lying in both cpco and cocp has only finitely many non-zero £ik, and conversely any vector with this property lies in both cpco and cocp; for this reason, we write cp co n co cp = q>, where we consider the coordinates of the vectors in cp to be indexed by pairs (i,k) of positive integers. Now let H1 be the linear subspace of cpco ©cocp consisting of all the (x,x) with atecpconcocp = cp. It is easy to see that H\ consists of all (at,—at), xecp, in cocp ©cpco. We denote this space by H2. H2=HU so that Hx and H2 are orthogonally closed both in cpco©cocp and in co cp© cpco. (1) H1 and H2 are orthogonally closed in cpco©cocp and HlnH2=o, but H1@H2 is not orthogonally closed. HX@H2 consists of all the finite vectors in cpco©cocp and so the closure of Hl@H2 is equal to cpco©cocp. This is in contrast to the behaviour of the closed subspaces of co (cf. § 9,3.(2)). (2) The quotient space ((pco®co(p)/Hl of the Uncomplete space L = cp co © co cp is not complete in the induced topology %k(L). Because H{ = H2 a cocp©cpco, it is enough, by §10,12.(6), to take the topology %k{H2) on ((pcjo + (jocp)/Hl. A linearly weakly compact
7. Further investigations 121 subspace of H2, however, is finite-dimensional. Given a sequence x{n) = (v){"\ 3(M)) of vectors in (pcoQaxp, the cosets x{n) with respect to H{ form a Xlk(H 2)-Oduchy sequence if the sequence is almost constant in each coordinate. In particular if we take the sequence / n \ x(n) _ / £ c_^ _ £ e ^ | tjien ^(«) js a cauchy sequence. But this has \i,k=i i,k=i J no limit in {cpa> ® a>cp)/Hx, for there is no x = (x)^)ecpoj®(D(p for which x — x{n) converges to 6; for suitably chosen (/,/), ^ = £^ = 0, and so if n^jj, we have (cj7, — ej7)(x —x(M))= —2. (3) Hj feas no closed complement in cpcoQaxp. For a complement would be ^-isomorphic to ((po)®o)(p)/Hl by § 10, 7.(6), and would therefore not be closed. The theorem on complementary spaces holds in cpco and cacp, however (cf. Hagemann [1]). The algebraic isomorphism H2 = H2/(H2nHl) = (H2 + Hl)/Hl is not an isomorphism for the topologies %lk(L). For H2 is closed, and so complete, whereas (H2 + Hl)/Hl is not complete, by (2). This is the counterexample mentioned in §10,7. 7. Further investigations. The theory of linearly topologized spaces has recently been developed systematically in works of Fischer and Gross [1], [2], [3], in order to provide a foundation for a theory of quadratic forms in infinite-dimensional vector spaces. A large part of the theory of linearly topologized spaces developed here applies when a skew field K is used as field of coefficients. In the definition of vector space over K given in § 7.1, it is necessary to restrict the definition in (L2) to x£; in this way we obtain a right vector- space or right module L over K. For the sake of simplicity, we have not developed the theory in full generality; cf. Dieudonne [6]. The circumstances are more complicated when K is supposed to be a general topological field. We refer to the investigations of Fleischer [1], Kothe [7], Nachbin [1] and Vilenkin [1]. The case of a non-archi- medean valued field has had detailed examination; it shows many similarities to the case of the real number field: cf. for example Bourbaki [6], Vol. 1, Fleischer [2], Ingleton [1] and Monna [1], [2]. Abelian groups, modules and rings with linear topologies have also been investigated; we refer to Ballier [1], Leptin [1] and Zelinsky[1]. If L is a linearly topologized space, if Ae2(L) and if C is a topological automorphism of L, CAC1 is said to be similar to A (cf. § 8, 7.). It is well known that, in the case where L is of finite dimension over the field V of complex numbers, the determination of canonical forms for A under "similarity-transformations" is achieved using the
122 § 13. The linear strong topology theory of elementary divisors. H. Ulm [1] has solved the similarity problem for a large class of endomorphisms of co(l~), using theorems about infinite abelian groups. The problem of the equivalence of quadratic forms can also be formulated in linearly topologized spaces. Let (L2,Lly be a dual pair. If A is a weakly continuous linear mapping from Lt into L2, so that Ae2(LuL2), then A' maps L2 = LX into L1=L2, so that A' is also in £(L1?L2). We call A symmetric when A — A. If A is symmetric, a symmetric bilinear form (y,x) -> y(Ax) is defined on LlxL1; the correspondence (x,x)^>x(Ax) is a quadratic form on Lv The function y(Ax) is continuous on Lx x Lx for the topologies Xls and Xlk, and every continuous symmetric bilinear form on L1xLl is determined by some Ae2(Ll9L2). If x = Cz is a topological automorphism of L1? an equivalent form z((CAC)z) is obtained from the quadratic form x(Ax)\ the problem then arises of classifying quadratic forms with respect to this equivalence. No general account of this problem has yet been given. For cp(P), co(P) and ^(P), however, it has been completely solved byK.Rnz- dorff [1].
CHAPTER THREE Topological Vector Spaces § 14 contains the elementary theory of normed spaces and Banach spaces. A number of classical examples are discussed, to which we shall refer time and again in the later parts of the book. The concept of topological vector space over the real or complex field is developed in full generality in § 15. In the investigation of these spaces we restrict our attention essentially to those questions which are of significance for the later theory of locally convex spaces. One important result is that every locally compact topological vector space is finite-dimensional, and indeed is topologically isomorphic to some P" with the usual topology. The spaces LP, with 0<p<l, are investigated to give an example to show that the dual space can consist of one element alone. Metrizable spaces are considered in detail. The classical results of the theory developed by Banach and his colleagues, together with a theorem of Bourbaki on bilinear mappings from metrizable spaces, bring § 15 to an end. In the next two paragraphs a detailed account is given of convex sets and the Hahn-Banach theorem. Provided that necessary care is taken, methods which go back to Minkowski can also be applied to convex sets in vector spaces of infinite dimension. We give three proofs of the Hahn-Banach theorem. The first two are of a geometrical nature, and are consequences of the separation theorem for convex sets, while the third is the classical analytic proof. The chapter ends with applications to normed spaces, and the theorem of F. Rmsz on the representation of the continuous linear functional on the space of continuous functions in terms of Stieltjes integrals. § 14. Normed spaces 1. Definition of a normed space. In this and the following chapters we shall only deal with real or complex vector spaces. From now on the coefficient field K means either the field P of real numbers or the field T of complex numbers. When it is not explicitly mentioned that we are dealing with a real or a complex vector space, then statements are always valid in both cases. A vector space E is called a normed space if a real number ||x||, the norm of x, is associated with each xeE, with the following properties: (Nl) |M|^0 for all xeE, (N2) if ||x||=0, then x=o,
124 § 14. Normed spaces (N3) ||Ax|| = |A|||x|| for each xeK, (N4) ||x + >>||^||x|| + M. If ||x|| only satisfies (Nl), (N3) and (N4), ||x|| is called a semi-norm on E. As well as ||x||, we shall use the notation p(x) or q(x) for a norm or semi-norm respectively. The simplest example of a normed space is K itself, with the modulus as norm. As in this case, we have in general that (1) |l|x||-|bl||^||x-y||^||x|| + ||y||. If we introduce a distance in E by putting (2) l*,J>l = ll*-J'll then it follows without difficulty that axioms (D1) to (D4) of § 4,1. are satisfied, so that every normed space is a metric space. The properties of metric spaces developed in § 4 are therefore also valid for normed spaces. The distance defined by (2) is translation-invariant, i.e. |x + z, y + z\ = \x,y\, for each zeE. The open ball Kr(x) of radius r about x, which consists of all yeE with \\y — x||<r, is obtained from the ball Kr(o) by a translation; Kr(x) = x + Kr(o). (3) The topology defined by the norm is compatible with the vector space operations, i.e. Xx and x + y are continuous in both variables together. If ||jc —jc0|| <^\\y-y0\\ <|, then,by(N4),||(x + ^)-(x0 + ^0)||<c so that x + y is continuous in both variables. It follows from \\xx-x0x0\\ = \\(x-x0)(x-x0)+{x-x0)x0+x0{x-x0)\\ ^|/-A0|||x-Xol| + |A-A0|||x0|| + |A0|||x-x0|| that we can make ||ax — /0x0||<c if we take \X — X0\ and ||x —x0|| to be sufficiently small. m m In particular it follows from (3) that £ 0$° xj^ converges to ^afcxfc if a[n)-+ak,x["^xk. The norm ||x|| is a uniformly continuous function on £, by (1). A linear subspace H of a normed space E is again a normed space, when the norm of E is restricted to H. If a semi-norm ||x|| is given on £, a normed space can be obtained in the following way: if ||x||=0 and ||y|| = 0, then it follows from (N3)
2. Norm isomorphism, equivalent norms 125 and (N4) respectively that |Ax||=0 and ||x + j/|| = 0. The elements of £ with vanishing semi-norm therefore form a linear subspace N of E. On the quotient space E/N let us set |x|| = ||x||, if x lies in the coset x. ||x|| is uniquely defined, since it follows from (N4) that ||x|| = ||x + z||, for zeN. All the properties of a norm are immediately satisfied by ||x||. Thus we have shown: (4) // ||x|| is a semi-norm on E and if N is the null-space of the semi- norm, then E/N is normed by \\x\\ = ||x||. 2. Norm isomorphism, equivalent norms. Two normed spces E and F are said to be norm isomorphic if there is a one-one linear mapping of E onto F, under which corresponding elements have equal norms. E and F are then isometric, as metric spaces. Two norms on the same vector space E are said to be equivalent if they define the same topology on E. (1) Two norms Pi(x) and p2(x) on E are equivalent if and only if there exist two positive numbers m and M such that (2) m ^ ^—^ ^ M holds for all non-zero x. Proof. Let K{r1](o) and K[2)(o) denote the open balls of radius r about o, with respect to the norms px and p2 respectively. If (2) holds, then it follows from the left hand inequality that K(rl)(o) is contained in X(r/2w\(o), and it follows from the right hand inequality that Xj.2)(o) is contained in Xj.^(o). The two systems of balls therefore form equivalent bases of neighbourhoods ofo. But since the metrics are translation-invariant it is enough to consider the neighbourhoods ofo, and so (2) is sufficient. Conversely, if px and p2 are equivalent the closed ball K{i\o) must contain a closed ball K{d2)(o). If p2(x) = S, it follows that p1(x)^l. We therefore have Sp1(x)^p2(x) for all x with p2(x) = 8. But this inequality must hold for all x, because of (N 3), so that the right hand side of (2) is established, with M=l/<5. The left hand side is established in a similar way. Two normed spaces E and F are said to be topologically isomorphic if there is an algebraic isomorphism of E onto F, which at the same time is a homeomorphism. If we have a topological isomorphism, and if we transfer the norm of E to F by giving corresponding elements the same norm, then we obtain a norm on F which is equivalent to the original one. (1) can therefore also be expressed in the following way:
126 § 14. Normed spaces (3) Two normed spaces E and F are topologically isomorphic if and only if there exist an algebraic isomorphism x1<-+x2 of E onto F and two positive numbers m and M, for which (4) m < < M 1 } ~ \\x2\\ ~ holds for,a\\ non-zero xx in E and corresponding x2 in F. 3. Banach spaces. A complete normed space is called a Banach space or (B)-space. (1) Every normed space can be embedded in a minimal (B)-space E, the completion of E, and this is unique up to norm isomorphism. The closed unit ball in E is the closure in E of either the closed or the open unit ball of E. The proof can be established using the corresponding theorem for metric spaces (§4,4.(1)). As a consequence of this, E is contained in a complete metric space E. If x and y are two elements of E, and if (x(M)) and (y{n)) are two sequences in E converging to x and y respectively, then |x,y| = lim|x(,l),j/(,l)|. If we put ||x|| = |o,x|, we obtain an extension of the norm from E to E, which clearly satisfies (Nl) and (N2). If we put Xx equal to the equivalence class of Cauchy sequences to which (Xx{n)) belongs, and likewise put x + y equal to the class to which (x(M) + y(n)) belongs, it is easy to see that unique elements of E are defined in this way (e. g. if z(M) -> x as well, then (Az(M)) and (lx(n)) are equivalent Cauchy sequences, since \\Xz{n)-Xx{n)\\ = \a\ \\z{n)-x{n)\\). Axioms (LI) and (L2) for a vector space, together with axioms (N3) and (N4) for the norm, are now established for E by taking limits in the corresponding axioms for E. Thus £ is a minimal (B)-space containing E. Two minimal (B)-spaces containing E are always isometric: because of continuity, sums and scalar products correspond under this isometry, and so the isometry is a norm isomorphism. Finally, if x04=o belongs to the closed unit ball of E and if xneE n—\ ||x0|| and x„^x0, then yn = -—-x„^x0 as well, since ||xj->||x0||^l, n ||xn|| and yn belongs to the open unit ball of E. Trivially, (2) A closed linear subspace of a (B)-space is again a (B)-space. It is often of interest to know if a (B)-space is separable. (3) // a normed space E is separable, then every subset, and in particular every linear subspace, is separable, and so is the completion E.
4. Quotient spaces and topological products 127 Proof. By §4,5.(1), separability is equivalent to the existence of a countable basis of open sets of E. This property is then satisfied by every subset. If, further, xt is a countable collection of elements which is dense in E, then xt is also dense in E. 4. Quotient spaces and topological products. Suppose that E is a normed space and that H is a closed linear subspace. The induced topology on the quotient space E/H is defined, as in §10,7., by the images K(0) of the open subsets O of E, where K is the canonical mapping of E onto E/H. (1) If H is a closed linear subspace of a normed space E, and if we introduce a norm on the quotient space E/H by putting (2) ||x|| = inf ||x|| (as x runs through all the elements of the coset x), xex then E/H becomes a normed space, whose topology is the induced quotient- space topology. Proof. If x#=o and x0ex, then inf ||xo + z||>0, for otherwise x0 zeH would be a closure point of the closed subspace H, and would therefore lie in H. Thus (Nl) and (N2) are satisfied. (N3) follows immediately from the corresponding (N3) for E, and (N4) follows from inf ||x + j/||^ inf ||x + j/||:S inf (||x|| + \\y\\) = inf||x|| + inf||j/||. x + yex + y xex,yey xex,yey xex yey Thus (2) defines a norm on E/H. If ||x||<p, then ||x||<p; conversely if ||x||<p, there is an xex with ||x||<p, and so the ball ||x||<p is the image of the ball ||x||<p under the canonical mapping. The topology defined by the norm (2) is therefore the quotient-space topology. (3) // E is a (B)-space, so also is E/H. Proof. We must show that E/H is complete. Preliminary remark: if x and y are two elements of E/H, and if xex is given, then, by (2), there exists yey with ||x — y\\ ^2||x — y\\. 00 First let xn be a Cauchy sequence in E/H with £ ||x„ — xn+1 \\ < oo. n= 1 Choose xx arbitrarily in the coset x1; using the preliminary remark, there is an x2ex2 with \\x{ — x2\\ ^ 2\\x1 — x2||, and, generally, there is an xM+1exM+1 with ||xM — xw_ x|| rg 2||xM — xll+1||. But it follows from 00 £ \\x„ — xM+1||<oo that xn is a Cauchy sequence in E. If x0 is its limit, n= 1 x„^x0 in E/H, since ||x„-x0||^ ||xM-x0||.
128 § 14. Normed spaces If xn is an arbitrary Cauchy sequence in E/H, there is a subsequence 00 x„k with Y, \\Xnk — xnk+l\\<co' This subsequence has a limit x0, and k= 1 this is also limit of the whole sequence. (4) // E is separable, so is E/H. We have seen in § 4, 7. that the topological product of countably many metric spaces can be given a metric which induces the same topology as the product topology. The topological product of arbitrarily many normed spaces is again a vector space, on which a topology is defined. We shall see in § 15,4. that even when there are countably many factors this topology cannot be defined by a norm. We can only assert: n (5) Let El,...,En be finitely many normed spaces, and let E= TT Et i= 1 be the topological product, with elements x = (x1,..., x„). If we define norms (6a) 11*11, = sup||x,||, (6b) IW|2=yil|x,.||2, (6c) ||x||3 = £||x,||, on E, then under each of these norms E is a normed space whose topology coincides with the product topology. The simple proof is left to the reader. There are other norms which have the same property. £ is a (B)-space if and only if all the Et are (B)-spaces (cf. § 4, 7.(2)). E is separable if and only if all the Et are. 5. The dual space. We consider the continuous linear functions with values in K on a normed space E over K. If ux is such a function, there must be a ball ||x|| <5 such that |wx|rg 1 for all x in the ball. For gen- d 1 eral non-zero z x = —z lies in the ball, and so \uz\ rg — ||z||. This gives one half of I'2'' " (1) A linear functional ux on the normed space E is continuous if and only if there exists M > 0 for which (2) M = M||x|| holds for all xeE. Since it is enough to establish continuity at o, (2) is clearly also sufficient. The collection of all the continuous linear functionals on a normed space E forms a vector space. We again call this the dual or conjugate space E of E. It is natural to introduce a norm on E' as well, by putting (3) ||w|| = sup \ux\. 11*11^1
6. Continuous linear mappings 129 It is trivial to show that this is a norm. The following inequality follows directly from (3): (4) liixl^M ||x||. (5) The dual E' of a normed space E is a (B)-space, when it is given the norm (3). Proof. If uin) is a Cauchy sequence, \\u(n) — w(m)||rgc for n,m^n0(e). By (4), \(u(n) — u{m))x\Ss\\x\l so that the sequence u{n)x converges for each x. We put vx = hmu{n)x. v is a linear functional on E. Since \(v — u(m))x\^e\\x\\ for m^n0, v — uim) is continuous. Since u{m) is also continuous, the sum v is, as well, i.e. v belongs to E'. Finally it follows from sup \(v — u{m))x\^£ that u{n) converges to r IMI^1 We are not yet in a position to show that there always exist non-zero continuous linear functionals on a (B)-space. The Hahn-Banach theorem (§ 17,6.) is required before we can do this. In this paragraph, however, we shall determine the duals of some classical examples of (B)-spaces. We can also form the dual of the space E'\ in this way we obtain the bidual space E" of E. The elements x0 of E can be interpreted as linear functionals x0(u) = ux0, ueE\ on E'. If the (B)-space £, considered in this way, is identical with the (B)-space E", E is said to be reflexive. Later in this paragraph we shall meet examples of both reflexive and non-reflexive (B)-spaces. If E is separable, E' need not be separable in the norm topology (cf.7.). 6. Continuous linear mappings. Let E and F be two normed spaces, and let A be a continuous linear mapping from E into F. Just as for linear functionals, we can establish (1) A linear mapping A from a normed space E into a normed space F is continuous if and only if there is an M > 0 for which (2) \\Ax\\^M\\x\\ holds, for all xeE. Let us recall that the continuity of a mapping between metric spaces can be defined in terms of sequential continuity, i.e. by requiring that x(«)^x(0) always implies that Ax{n)^Ax{0\ The vector space 2{E,F) of continuous linear mappings from E into F is turned into a normed space, when we put (3) 11-411 = sup \\Ax\\. 9 Kothe, Topological Vector Spaces I
130 § 14. Normed spaces We observe that (4) Mxll^MHW. (5) // F is a (B)-space, then 2(E,F) is a (B)-space. The proof of 5.(5) carries over directly. The topology defined on 2(E,F) by (3) is called the uniform norm topology. (6) // A is a continuous linear mapping of E into F, and if B is a continuous linear mapping of F into G, then the continuous mapping BA of E into G satisfies (7) ||B/t||g||B||M||. For by (4), \\BAx\\ £ \\B\\ \\Ax\\ g ||B|| \\A\\ \\x\\, so that sup \\BAx\\ £\\B\\\\A\\. M = > If a normed space E is an algebra over K in the sense defined in § 8, 1., and if the product xy of two elements of E always satisfies (8) ||xj;||^||x||||j;||, then E is called a normed algebra. If, further, E is a (B)-space, then we call it a Banach algebra. It follows from (7) that (9) The set 2(E) of continuous endomorphisms of a normed space forms a normed algebra with unit element. If E is a Banach space, then 2(E) is a Banach algebra. It follows from (8) that \\xy-xnyn\\ ^ \\x\\ • \\y-yn\\ + ||j/J • ||x-xj, so that the product xy is continuous in both variables together; condition (8) therefore means that the product operation is compatible with the normed space topology. This idea also applies to sequences of mappings Ane2(E,F), Bne2(F,G): if An^A and Bn^B, then BnAn^BA in £(£,G). The theory of Banach algebras will not be considered in this book; we refer the reader to Hille and Phillips [1], Loomis [1] and Naimark [1]. 7. The spaces c*0, c, I1 and /°°. If we interpret a bounded sequence £fc, k= 1,2,..., of real or complex numbers as a coordinate vector x = (£k)9 and if we define -e + n to be the vector with coordinates £k-\-rjk, where r) = (rjk), and define Ax to be (A^), then we obtain a real or complex vector space /°°. We introduce a norm on /°° by setting (1) Na. = IWI = sup|£j.
7. The spaces c0,c,ll and /' 131 The norm properties are satisfied. Moreover, (2) Z00 is a (B)-space. For if x{n) = (&n)) is a Cauchy sequence in /°°, then sup l^-^l^e k for all n,m^n0. Thus for fixed k each coordinate sequence £{kn) converges to some £{k°\ so that x{n) is coordinatewise convergent to x{0) = (£[0)). Further, sup |^M)-^0)|^e, for n^n0, so that x{n)^xi0\ and x(0) is a k bounded sequence with \\x{n) —x(0)|| ^ c; i.e. x(M) has x(0) as limit in /°°. We denote by c and c0 the linear subspace of /°° consisting of all convergent sequences and all sequences convergent to 0 respectively. We give both spaces the norm (1). Then we have (3) c and c0 are (B)-spaces, and so are closed linear subspaces of /°°. Proof. Let x{n) = (£kn)) be a Cauchy sequence in c. We must show that its limit x(0) = (^(fc0))= Uim ^"M, which exists by (2), is again a convergent sequence. By hypothesis \\m^kn) = ^{n) exists, for each n. Given s>0, there k-* oo exists an n0 such that sup \^kn) — ^km)\ < — for n,m^n0, so that k 3 for sufficiently large/c, and for n,m^n0. Thus lim^(M) = ^(0) exists. n-* oo We now show that lim 40) = £(0)- Given £ > 0, let n0 be large enough k-* oo for |£<">-£«»| ^ - and sup|^",-4°,| £ - to hold for n^n0. If for a 3 * 3 e fixed n1^.n0 we choose k0 large enough for l^"1' —^"''l ^ — to hold for all fc^fc0) then 3 \£0)-?0^\&0,-&Hl)\ + \&Hl)-Pin + \?mi)-?0)\£s for k^k0, as required. The proof for c0 is included in this. 00 We denote by Z1 the vector space of all x = (£k) with ]T 1^1 <oo. This is normed by k=1 J 00 (4) IWI1 = W=EI^|. k=\ (5) Z1 is a (B)-space. The simple proof is left to the reader. It is also contained in the proof of 8. (7). 9*
132 § 14. Normed spaces If we only consider the vector space structure, we clearly have (6) Z1 c=c0c=c(=/ao. The question of separability is easily answered: (7) Z1, c0 and c are separable, and /°° is not. Proof. The vectors with finitely many non-zero rational coordinates are dense both in Z1 and c0. The vectors with finitely many arbitrary rational coordinates and with the rest equal to one fixed rational are dense in c. In /°° two vectors whose coordinates are equal to +1 or — 1 are always distance 2 apart. Since there is a continuum of these, the set of these vectors is not separable, and so by 3.(3) neither is /°°. We now attempt to determine the dual spaces precisely. (8) r is the dual of I1. Proof. More precisely, our assertion says that (Z1)' is norm isomorphic to /°°, so that each continuous linear functional u(x) can be represented, in a way yet to be defined, by a vector u in /°°. We shall come across this problem of the concrete representation of abstract linear functional more and more often. We shall formulate the line of thought in a rather more general setting, as the same idea can be applied in other situations. A set M of elements of a normed space E is called a fundamental set if its linear span is dense in E. We also say that M is total in E. A continuous linear functional is clearly uniquely determined by its values on a fundamental set. The unit vectors ep, p= 1,2,..., form a fundamental set in Z1, so that if we(/1)', u is determined by the vector u = (vp\ where vp = u(ep). If x = {£p)el\ it follows from the fact that the sections xM = ((^1,...,^„,0,0,...) converge to x that n oo (9) u(x) = lim u{xn) = lim £ vp tp = £ vp £p = u x. n n l l If we put x{p) = cplep when v = sp\vp\+09 and put x{p) = ep if vp = 09 then ||*(p)|| = l, so that ux{p)= \vp\ ^ ||u||; i.e. u lies in /°°, and in /°° it has norm ^ ||w||. On the other hand, oo) i«(*)i^(suP|0,i) £ ig, p p= i
7. The spaces c0,c,ll and /' 133 so that ||w|| ^ ||u||. It therefore follows that ||w|| = ||u||, which gives the required norm isomorphism. It follows immediately from (10) that conversely every ue/°° defines a continuous linear functional on Z1. We observe that the dual of the separable space Z1 is not separable. (11) I1 is the dual space of c0. The sections xn converge to x in c0 as well, and we again obtain n u(x) = \imYJvp£p = ux, for a continuous linear functional w, where vp = u(<tp). Again,let vp = t:p\vp\ n if vp + 0, let sp=l if vp = 0, and let x' = X£p1er Tnen ll^'ll = 1 m co> » i so that ux' = £ \vp\ ^ ||w||. Henceu lies in Z1, and ||u||^||w||. Conversely, \u(x)\^\vp\\\x\\, so that ||u|| ^ ||u||. 1 (8) and (11) show that the bidual of c0 is equal to Z00, and is therefore larger than c0. Hence c0 is an example of a non-reflexive (B)-space. We now consider the dual of c. Here we also find that c' = /1, although the representation is a different one. The space c contains c0 as a subspace, and c0 has codimension 1. (12) c = co0[e], where [e] is the one-dimensional space consisting of the scalar multiples of the vector e, which has all its coordinates equal to 1. Every convergent sequence x = {£„) can be represented in the form (13) x = £0e + x0 with 10 = lim tn and x0ec0. n-* oc A fundamental set in c is formed by e and the vectors ep, p= 1,2,.... If u(x) is a continuous linear functional on c/let us put i/0 = w(e) and vp = u(ep), for p^l. It follows from (13) and (11) that 00 u(x) = u(£0*) + u{x0) = £0Vo + X>p(£p-£o) with (ui,^,--.)^/1. 1 00 If we put u = (i;0,t?1,...), where v0 = v'0-Y,vp9 and put x' = (€0,€i, •••) then i (00 \ 00
134 § 14. Normed spaces Conversely, if uel1, (14) always determines a continuous linear functional with ||w||<||u||, since lus'i^ XKI supigniuiisupigHluiuisL. The inequality ||u||^ ||w||, and hence the norm isomorphism, is established in jthe following way. Again let vp = ep\vp\ for vp + 0, and sp=l for vp = 0, and let x{n) be the convergent sequence, the first n of whose terms are equal to ej"1,..., e~l, and whose subsequent coordinates are n all equal to e^1. Then ||x(")||ao = l, and \u(x{n))\ = \ux{n)'\ ^ \v0\ + £ \VP\ i oo n - X |wp|. It therefore follows from \u(x{n))\ ^ ||u|| that Kl + X>pl M + 1 1 00 - Y, KJ^IMI- Letting n^oo, this gives ||u||^||w||. M+l We therefore obtain (15) The dual space of c is norm isomorphic to Z1, when we make the continuous linear functional u correspond to the vector u = (v0,v1,...), 00 where vp = u(ep) for p=l,2,... and v0 = u(z) — Yvp> an^ wnen we apply u to x in the way defined by (13) and (14). * c is therefore not reflexive either. There is no such simple representation of the dual of Z°°, but we shall see later on that Z1 and /°° are also not reflexive. We observe that at any rate Z1 is a linear subspace of (/°°)', when we interpret each ue/1 as the linear functional ui on /°°. A detailed investigation of (/°°)' is made in § 31, No. 1. 8. The spaces lp91 <p< oo. For the sake of completeness, we prove the fundamental inequalities upon which the theory of these spaces depends. (1) // 0<a<l and a^O, 6^0, then (2) aaZ?1_a^aa + (l-a)Z?. Proof. If a = b we have equality. We can clearly suppose that b > a > 0. By the mean-value theorem, bl-*-al-* = (l-a)(b-a)Z-* with a<^<b. Since £~a<a~a, we have b1~a-al-a<(l-a){b-a)a-a. Multiplying by aa, we obtain (2).
8. The spaces lp, 1 < p < oo 135 We remark that the proof shows that equality only holds in (2) when a = b. Using this result, we prove Holder's inequality oo /oo \ l/q /oo \ 1/p J ] (3) 5>*&l^(2>J'J (£^|p) 7 + ^" = 1' 1<p<<x>- Here yfc and ^ are arbitrary real or complex numbers for which the sums on the right-hand side converge. n cc It is clearly sufficient to establish (3) for £ instead of for £. 1 1 If we put a = —, l-a = —, a = cpk, fc = <#, in (2), we obtain P 4 (4) Cfcdfcg_cE + _d£. P 4 Letc* = 7^ yTP>dk = 7~n yTq for fc=1> •••>"• Summing over (l\Si\p) [Ih\q) the resulting inequalities, we obtain n " v/*/n v/p"p Zl^lp + <? Zkl* y/q/n \1 which establishes (3), with n in place of oo. We remark that equality holds in (3) if and only if it holds in all the inequalities (4), and so if and only if one of the two vectors {\vk\q) and (\£k\p) is a scalar multiple of the other. From Holder's inequality we obtain Minkowski's inequality /oo \l/p /oo \l/p /oo \l/p (5) (ll^ + ^r) ^(ll&l'J +(Zl^lPj > l<P<oo. Proof. We have Zi^+^i^Zi^ii^+^' + ZKii^+^r1- i i i Applying (3) to each of the two summands on the right-hand side, and bearing in mind that (p—\)q = p, we obtain =[(Zl4lp)1/p+(ZWp),/p](Zl^+^lp)1M-
136 § 14. Normed spaces If we bring the last factor across to the left-hand side, then because 1 = — the left hand side becomes (Y \£k + rik\p)1,p, so that we ob- q p tain (5), with n in place of oo. Again, equality holds in (5) if and only if one of the vectors (£k) and (rjk) is a non-negative scalar multiple of the other. We denote by /p, 1 <p< oo, the collection of real or complex vectors 00 i = (^1,^2j-.-) for which £ \£k\p converges. The expression k=1 /oo \i/P (6) 11*11 = 11*11,= £K is used to define a norm on lp. The norm properties are trivial except for ||s + t)|| ^ ||*|| + ||t)||, and this is Minkowski's inequality. (7) /p, 1 <p< oo, is a (B)-space. Proof. For a Cauchy sequence x{n\ Xl^n)-^m)|pgep holds for i n,m^n0, and this implies the coordinatewise convergence of x(n) to some x<0> = (<*°>). It also implies that £ \Zktt)-Zk0)\p^ep for each r, so that 00 1 El&n)-&0)lP^fiP- But this means that x(n)-x(0) lies in /p, and there- i fore that xi0) does too, and that x{0) is the limit of the sequence x(n). (8) Each /p, l<p<oo, is separable, and the dual of lp is lq, where 1 1 - + -= 1. p q. Proof. The unit vectors ck, fc=l,2,..., again form a fundamental set, and their rational linear combinations are dense in /p, so that lp is separable. From the fact that the sections xn of x converge to x it follows once again that a continuous linear functional u can be represented by n oo u(x) = lim £ vk£k = X i>fc£fc = uat with ufc = u(efc). n-oo x x Let ufc = efc|uk| for |ufc| + 0, let efc=l for ufc = 0, and put i The norm of x{n) in /p is equal to l/p /n \l/p ¥% = (tMq-1)pJP=(t\^
9. (B)-spaces of continuous and holomorphic functions 137 Hence |u(x<">)| = £ \vk\>£ \\u\\ (£ MJ, so that (zw)1"' = (|:ki,y/^iNu i. e. u lies in lq and ||u||^||w||. If, conversely, uelq, then because of Holder's inequality we have M|^X>^|^||u||j*||p, 1 so that u defines a continuous linear functional u on lp, and ||w|| ^ \\u\\q. I2 is Hilbert space, which is therefore dual to itself. By (8), all the spaces lp, 1 <p< oo, are reflexive, in contrast to I1 and /°°. (9) // l^/?i<P2=°°> JP1 *5 fl proper subspace of lP2. The topology on lPl defined by lPl is coarser than the norm topology of lPi, and \\x\\P2 ^\\*\\Pl, M each xelp\ 00 Proof. If at is an element of the unit sphere of lp\ so that £ |£fc|pi = 1, oo ! then Xl^fclP2=l f°r Pi<P2<co> and sup |<^fc|^ 1 for /?2 = °o. It i k follows that ||x||P2^||x||pl for these xelpi. But this inequality also holds for all Xx, and it therefore holds for all xelpi. Pi If p = , and if Pi<p<p2 then clearly the sequence £>k=k p, 1+e /c=l,2,..., is an element of P2 which is not in lpi. The spaces lp (l^p^co) can also be defined for general index sets. If a runs through an index set A with cardinal d, then, for p<co, let x = (£a) denote a vector with d coordinates, of which at most countably many are non-zero. Then (6) has a meaning when the sum is taken over all a; lp is the normed space of all those x for which ||x||p is finite, and ||x||p is again the norm. Once again lp is a (B)-space with dual /jj, for 1 <p< oo. The proof we have given holds in this case as well. The dual of/] is equal to If, which consists of all the vectors with ||x|| ^ = sup |<? J < oo. If d is uncountable, the spaces lp are no longer separable. If d is finite, so that d = n, we obtain finite-dimensional normed spaces. The dual of If is /*, as can be seen immediately; If is therefore reflexive. 9. (B)-spaces of continuous and holomorphic functions. Let K be a compact space. We have shown in § 6, 4. that there are "sufficiently many" continuous functions defined on the whole of K (both real- and complex- valued), and by § 6,2.(7) sup \f(x)\ is always finite. We have xeK
138 § 14. Normed spaces (1) The vector space C(K) of all real-(respectively complex-) valued continuous functions defined on the compact space K is a (B)-space when we give it the norm (2) ||/|| = sup |/(x)|. xeK Proof. The linear operations are defined in the usual way: (fi+f2)(x)=Mx)+f2(x),(Xf)(x) = />(f(x)). If f„ is a Cauchy sequence in C(K\ there is for each e>0 an n0 with \fn{x)—fm(x)\ ^— for n,m^n0, and for all xeK. The existence of a limit function f0(x) follows from this, and \fn(x)—f0(x)\ ^ — for n^n0 and for all xeK; i.e. the/„ converge to f0(x) uniformly on K. Finally f0(x) is continuous, for there is a neighbourhood U(x0) in which \fn(x0)—fn{x)\ ^ —, so that l/oM-/o(*o)l = l/oW-/nWI + l/-M-/n(^o)l + l/n(*o)-/o(*o)l ^ for all xeU(x0). The dual of C(K) is called the space $ft(K) of measures on K. We shall study this in detail in the second volume. We shall consider the classical result of F. Riesz in the case where K = I = \_0,1] as an example, in §17,7. C(I) is separable. In the case where the functions are real-valued this follows directly from the Weierstrass approximation theorem, since any continuous function can clearly also be approximated uniformly by polynomials with rational coefficients. The complex-valued case is then trivial. Besides these spaces, we shall have to consider spaces of holomorphic functions. First let © be a bounded open domain of arbitrary connectedness in the complex plane. Suppose that the closure © of © consists of © together with a system C of finitely or countably many boundary curves Cr We consider the functions f(z) defined on ©, which are holomorphic in © and continuous on ©. They clearly form a vector space HJ5(©), and we have (3) // we introduce a norm on HB((&) by setting (4) ||/|| = sup \f(z)\, ze(5 HB((&) becomes a (B)-space. For if /„ is a Cauchy sequence with respect to the_ norm (4), fn is uniformly convergent to a continuous function f0 on ©, which is also holomorphic throughout ©, by a theorem of Weierstrass.
10. The//-spaces (p^l) 139 We shall also consider the case where © consists of finitely many open domains ©£ in the complex sphere Q, whose closures ©^ are disjoint. © must be different from Q itself. A function f(z) defined on © is said to be locally holomorphic in © if it is differentiable at each point ze®, and if it vanishes at the point oo, if oo belongs to ©. Now let HJ5(©) be the space of functions defined on © which are locally holomorphic in © and are continuous on ©, with the norm (4). (3) also holds for this space, and it can be seen without difficulty that it is the topological product of the HJ5(©t), with the norm (6 a) of 4. 10. The L^spaces^^l). If, for a continuous function f{x) on / = [a, b] (— oo < a < b < + oo), we take the expression (1) \\f(t)\pdt l^/7<00, as a norm, then the norm properties are satisfied, as we shall see presently; it is easily established however that a Cauchy sequence in this normed space of continuous functions need not have a limit. By 3., this normed space which we have just defined has a completion. We denote this completion by LP(I\ or LP for short. Rather than being satisfied with this abstract definition, however, we give a concrete representation of LP, using the theory of Lebesgue measure and integration. Here we assume that this theory is known, although general measure theory will be developed in the second volume, and this contains the classical theory as a special case. The elementary LP-spaces considered here will also appear as special cases of general LP- spaces. A real- or complex-valued measurable function f(t) defined on 7 = [a,fr] except perhaps on a set of measure zero, is called p-th power summable (or integrable) (p^l) if (2) Jl/(0lp<"<oo- a For fixed p ^ 1 let Lip) be the set of all measurable p-th power summable functions on /. For each feL{p\ (1) takes a finite value. ||/||p = 0 if and only if/ vanishes except on a set of measure zero. (3) L(p) is a vector space. Clearly kf is in Lip) iff is, and ||A/||P= W \\f\\p. If a and b are arbitrary complex numbers,and p^l, (4) \a + b\p^{2m^x{\a\9\b\)p^2p(\a\p+\b\p).
140 § 14. Normed spaces It follows from this that j \f+g\pdt^2p{\ \f\pdt + j \g\pdt)\ i.e. f+g is in L(p) if/ and g are. We again prove Holder's inequality: (5) If f{t)eLip\g(t)eLiq\- + - = l,p>l, then f(t)g(t) belongs to to L(1), and P 4 (6) fl/Wff(t)|dt^(jV(Olpdt) "• (j\g(t)\qdtjlq= \\f\\p\\g\\q. Proof. If ||/||p = 0 or ||gf||^ = 0, then the left hand side is also zero, so that the inequality holds. Otherwise we can put c = ——— and d = ——- in 8.(4), to obtain HA ^ (7) l/^l 1 \f\P 1 lglg ll/IUIffll,= P 11/115 9 llffllj" /gf is measurable, and the absolute value of fg is majorized by an integrable function, by (7), so that fg is integrable; i.e. fg lies in L(1). Integration of (7) gives (6). (8) Minkowski's inequality. If /, geLip\ p^l, then (9) ii/+ffiip^imi,+yip. For p=l the assertion is trivial. For p>\ it again follows from Holder's inequality (which can be applied, since f+g lies in Lip) by (3), so that If+gl"'1 lies in L{q\ as (p-l)q = p): l\f+g\Pdt^\f\\f+g\p-'dt + \\g\\f+grUt ^(\\f\\P+\\g\Mf+o\Pdt)1/q- Again we observe that for p > 1 equality holds in (6) if and only if one of the functions \f\p and \g\q is a scalar multiple of the other, almost everywhere, while equality holds in (9) if and only if one the functions / and g is a non-negative scalar multiple of the other, almost everywhere. Thus we have shown that in L(p) the expression (1) satisfies all the properties of a norm with the exception of (N2). Thus (1) only defines a semi-norm on L(p). The null-space of the semi-norm is the space of all those functions which vanish almost everywhere. By 1.(4), Lip)/N is a normed space, under the norm (1). It is customary to make no distinction between feLip) and the equivalence class in Lip)/N determined by / It is readily confirmed that all the results for functions in Lip) contained in this number also hold
10. The //-spaces 141 for the equivalence classes in Lip)/N, since these consist of functions which are equal almost everywhere. (10) Lip)/N is a {Byspace. Proof. Let /„ be a Cauchy sequence in Lip)/N. We say that the functions fn form a Cauchy sequence in the p-th mean. Then for each v there is an nv>nv_1 for which \\fm—f„\\p < — holds, for m,n^rcv. In particular, 1 y' (ii) ll/„v+1-/„J5< If now Mv is the set of points for which |/„v + 1(0-/nv(0l>2"v/p, then the measure m(Mv)<(f)v, for it follows from (11) that 2"vm(Mv)^ \fnv+l-UPdt<jv. If t does not lie in (J Mv, then |/nv + 1(t)-/nv(t)|^2"v/p for v^JV + 1, oo N+l oo so that X l/-v+ M-LM <«d for these t; £ (f^.tf-fjt)) there- v= 1 00 00 fore converges for all t which do not lie in M = f] \J Mv. The set M N=\ N+l has measure zero, and limfn (t) = f(t) throughout I~M. On M we put/(t) = 0. We now show that f(t)eLip\ and that /„ converges to / in p-th b mean. j\f„v(t) —fn{t)\pdt^s, for sufficiently large nv and n. The func- a tions |/f,v(0—/ii(0lp are non-negative, and as wv-»oo they converge almost everywhere to \f{t)-fn{t)\p. ApplyingFatou's lemma, \f(t)—f„(t)\p is integrable, and $\f(t)-fMpdtSiim$\fnv(t)-fn(t)\pdtSc. a a Thus f—f„, and consequently /, lie in Lip\ and ||/—/Jp->0. We now establish (12) L{p)/N is identical with U. The space of continuous functions on I = \_a,b~\ can be considered as a subspace of Lip)/N, for two distinct continuous functions never differ by a function in N, so that the embedding into Lip)/N is one-one.
142 § 14. Normed spaces If we can show that the continuous functions are dense in Lip)/N, it will follow from (10) and the uniqueness of the completion that Lip)/N and U are identical. Since every real valued function in Lip) can be expressed as the difference of two non-negative functions in Lip\ and since every complex- valued function can be expressed as fi—f2 + i(f3-/4)*fi = ®> ^ *s enough to show that a non-negative / can be approximated in p-th mean by a continuous function. As is shown in the theory of real functions, given e > 0, 3 > 0 and a non-negative measurable function f(t) on /, a continuous function cp(t) can be found for which \f(t) — q>(t)\<6 for all t outside some set of measure <£. If f(t)^K, cp{t) can also be chosen with cp(t)^K. b If now 0^f{t)^K then $\f(t)-<p(t)\pdt^8p(b-a) + Kpe9 so that a f can be approximated by continuous functions. If f(t) is unbounded, we introduce the functions fn{t\ which are defined by f„(t) = f(t) if f(t)^n, and /„(*) = 0 for f(t)>n. If M„ is the set on which /„ vanishes, then \\f—fn\\= I J \f{t)\pdt) < — for suf- M n ficiently large n. Since a continuous function cp can be chosen for which 8 \\f„ — (p\\ < —, the approximation property also holds for unbounded feLip)/N. 2 (13) L!\ p^\, is separable. For it follows directly from the Weierstrass approximation theorem that C[7] is separable under the norm (1); so therefore is its completion LP. We remark without proof that I2 and L2 are norm isomorphic. This is essentially the Riesz-Fischer theorem. Analogous to the results about the V spaces, we have (14) The dual space of LP is Lq, where — + — = 1 and 1 <p, q< 00. P 4 We shall give the proof later (§ 26, 7.), using more general ideas. The space L00. A measurable function f(t) defined on I = [a,b\ except perhaps on a set of measure zero, is said to be essentially bounded if there is an M^O for which \f(t)\^M for all t outside a set of measure zero. By the essential supremum of such a function, in symbols ess sup| f(t) |, we mean the greatest lower bound M0 of all the essential upper bounds M. M0 is itself an essential upper bound,
The space L^ 143 for if \f(t)\^Mh f=l,2,...,M1>M2> •••-^M0, and if Nt are the sets 00 on which |/(r)|>Mf, then N0 = (J iVt- is again of measure zero. i= 1 We denote by L°°(7) the vector space of all essentially bounded measurable functions on / where again we identify functions which are equal almost everywhere. We introduce a norm on L°°(7) by putting (1) ll/IL = esssup|/(t)|. tel For this we have (2) La is a (Byspace under the norm (1). The norm properties are trivially satisfied. Let /„ be a Cauchy 1 1 sequence. The inequality ||/n-/mL^- means that \f„(t)-fm(t)\^- for all t outside some set N{®m of measure zero. The sequence fn(t) is 00 therefore uniformly convergent for all t outside the set N0= (J N(^m m,n,k= 1 of measure zero. The limit function f0(t) defined on I ~N0 is measurable and bounded, and so lies in L00. It is clearly the limit of the sequence fn. The'(B)-space C(I) with the norm defined in 9.(2) is a proper closed linear subspace of L00, and so C(I) is not dense in L00. (3) L° is not separable. If we divide / = [0,1] into a sequence of intervals /„ = , — \_n + l n and consider functions f(t) which take a fixed value +1 or — 1 on each In, then we obtain uncountably many functions on L00 any two of which are distance 2 apart. The inclusion relations between the various LP spaces are the exact opposite of the relations between the lp spaces. (4) // 1 ^p1 <p2^ °°> LP2 is a proper subspace of LP1. The topology induced on LP2 by LP1 is coarser than the norm topology of LP2, and if fell2 we always have (5) ll/llP1^ll/Ub-a)1/p,-1/P2, l^p,<p2^oD. IffeL", ||/IU=lim||/||p. p->oo Proof. Let feLP2, p2<oo. For those t for which |/(0I^1 we have l/Wr^l/WT2, so that |/(t)|pi is majorized by the integrable function #(x) = max {1,|/(t)|P2}; thus |/(r)|Pl is itself integrable and feLP1. If p2 = oo, it is trivial that L00 c LP\
144 § 15. Topological vector spaces If p2 = oo, (5) is also trivial. Let p2<oo. We apply Holder's inequality to \f\PleLPllp\ g=l and the exponent p = p2/Pi to obtain b /b \pi/P2 P2-pi lurdtz[Wfrdt] (b-a) ** a \a J from which (5) follows. Finally let ||/||oo = M>0. Then given e>0, there is a set N of positive measure \i > 0 on which \f(t) | ^ M — e. Hence \\f\\p ^ \_ii(M — e)p~\1/p = //1/p(M-e), so that lim ||/||p^M; it follows from (5) that H/L = lim ll/ll,. P~* OO If we take the underlying interval to be / = (—00,00), then the definition of LP( — 00,00) must be made differently. In this case we start from the space C00(—00,00) of all continuous functions on / which vanish outside a finite interval. For these functions /+00 \i/P (6) ll/llp = (^ I \f(t)p\dtj is finite, and we define LP( — 00,00), 1 ^/?< 00, to be the completion of the space C^ with the norm (6). On the other hand the definition of L(p) and L00 is again the same as for a finite interval [a,b], and the preceding results and proofs remain unaltered, up to the following point: The assertion of 10.(12) still holds, so that Lip)/N is equal to Lp, but the proof must be changed. We show that C^ is dense in Lip)/N. Because of the convergence of (6), every feLip) is approximable in /?-th mean by an feL{p) which vanishes outside some finite interval. Using the proof of 10.(12), this / is approximable by a continuous function vanishing outside the same interval, so that C^ is dense in Lip)/N; i.e. Lip)/N = Lp. The separability of LP again follows from the separability of Cx. The space of all bounded continuous functions on ( — 00,00) is a proper closed linear subspace of L00. Theorem (4) of this number is false for LP( — 00,00); indeed, L00 is not contained in any of the spaces LP. (7) The dual space of L1 is L00. This theorem will also not be proved until later (cf.§ 26,7.). § 15. Topological vector spaces 1. Definition of a topological vector space. In this paragraph we consider vector spaces over the field K of real or complex numbers; this is given its usual topology by means of the modulus.
1. Definition of a topological vector space 145 The ideas are closely related to those of § 10.2., where of course we were concerned with an arbitrary field, with the discrete topology. A vector space E over K is called a topological vector space or a topological linear space E[X~\ if a Hausdorff topology Z is defined on E which is compatible with the vector space structure. Compatibility once more means that the mappings (x,y)^>x + y from ExE into E and (a,x)->ax from KxE into E are continuous. The concept goes back to A. Kolmogoroff [1] and J. von Neumann [1]. The topological isomorphism of two topological vector spaces EifXi] and £2[^2j> m symbols E[^X1]^E2[^2]? *s defined as in § 10,2. as a vector space isomorphism which at the same time is a homeomorphism. By § 14,1.(3), a normed space is an example of a topological vector space. Giving a norm, however, provides a richer structure than that provided by the topology which it induces; as we saw in § 14,2., two normed spaces can very well be topologically isomorphic as topological vector spaces without being norm isomorphic. If follows directly from the definition of topological vector space m m that X!afc,)x*n) converges to Yjockxk with respect to X if each of the i i sequences x^\ n=l,2,..., converges to xk with respect to £ and each sequence o4n)->afc in K. (1) The mapping x->x + x0 is a homeomorphism of E[X] onto itself. The mapping x->ax, a4=0, is a topological automorphism of E[%~]. For these mappings are continuous, and linear and continuous, respectively, and have the whole of E[X] as image space; the inverse mappings x^x — x0 and x-► — x exist and have the same properties. a If U = {U} is a base of neighbourhoods of o, the sets x0 + U therefore form a base for the neighbourhood filter of x0. Further, if U is a neighbourhood of o, so is olU, for a 4=0. (2) // E\%\ is a topological vector space over K and U={U} is a base of neighbourhoods of o, then we have (LT1) For each UeU there is a VeU with V+V^U. (LT2) For each UeU there is a VeU for which olV^U for all oc with |a|rg 1. (LT3) For each UeU and each xeE there is a positive integer n(x,U) for which xenU. A subset M of E is said to be absorbent if a suitable multiple px, p > 0, of each element x of E lies in M; (LT 3) says that every neighbourhood of o is absorbent. 10 Kothe, Topological Vector Spaces I
146 § 15. Topological vector spaces Proof of (2). (LT1) is nothing but the continuity of (x,y)->(x + y) at (o,o). It follows from the continuity of ax at (0,o) that there is an 8>0 and a neighbourhood Wofo such that £xeU for all xeW and all |^| ^6, so that (LT2) is satisfied with V=s W. If U were not absorbent, there would be an x0eE which would not lie in any n U. Thus — x0$ U would hold, which contradicts the conver- •1 n gence of — x0 too. n A set Mc£ is said to be circled if ax0 belongs to M whenever |a|:gl and x0eM. The circled cover of a set M consists of all ax, xeM, |a|^l. (3) A topological vector space E always has a base of neighbourhoods of o consisting of circled neighbourhoods. For the sets [j ocU, UeU, form a base of neighbourhoods of o, by(LT2). i«|^i 2. A second definition. We now establish the converse of Theorem (2) of the preceding number. (1) Let VL = {U} be a filter base on a real or complex vector space E, with Q U = o, which satisfies conditions (LT 1) to (LT3). // a topology X is defined on E by taking as neighbourhoods the sets U(x) = x+U, UeVL, then E[X~\ is a topological vector space, with U as base of neighbourhoods of o. By 1.(2), every topological vector space can be defined in this way, for by the remark after 1.(1) the topology on E is always determined by a base of neighbourhoods ofo. The following definition is even simpler: (2) Suppose that VL = {U} is a filter base of absorbent circled sets U on a real or complex vector space E; suppose that f]U =o, and that for each U there exists a VeU with V+ Va U. If we define a topology X on E by taking as neighbourhoods the set U(x) = x+U, UeU, then E\X~\ is a topological vector space, with H as base of neighbourhoods of o. Once again, every topological vector space can be defined in this way, by 1.(2) and 1.(3). It is sufficient to establish (2), for, given a filter base H which satisfies the hypotheses of (1), we obtain an equivalent filter base satisfying the hypotheses of (2), just as in the proof of 1.(3). Proof of (2). We define the open sets of E to be those sets which contain a neighbourhood U(x) = x+U of each of their points x. It is easy to confirm that the class O defined in this way satisfies axioms (01) and (O 2) of § 1,1., so that we have indeed defined a topology X on E. X is Hausdorff: suppose that x#=y. Since f] U = o there is a U with
2. A second definition 147 x-y$U. If V+VcU, with V circled, then x+V and y+V are disjoint. For if x + z1=y + z2, with zt,z2eV, then x — y = z2 — zleV + Fc[/, contrary to our assumption. The continuity of the mapping (x,j/)->x + j/ at (x0,y0) follows from (x0+V) + (y0+V)czx0 + y0+U. The continuity of the mapping (a,x)-^ax at (a0,x0) is shown as follows: suppose that we are given a neighbourhood a0x0 + C/, with U circled, and suppose that |a0|^n. We can then find a circled V such that V + ••• + V (n + 2 summands) lies in U (this must be distinguished from (n + 2)V, which only contains the elements (n + 2)x, xeV). A fortiori, nV+V+Va U. Further, let us choose a positive integer m such that x0emV. . If now |a — a0| ^ —, xex0 + V, the relation otxeot0x0 + V+V+nV m cza0x0 + U follows from ax = a0x0 + (a — a0)x0 + (a — a0)(x — x0) + a0(x — x0) 1 1 since — (mV)=V, (a — a0)(x — x0) e— Kc V and oc0(x — x0)enV. Hence m m ax is continuous. (3) Every topological vector space £[JX] is uniformizable, and is therefore regular (and indeed, by § 6, 8.(1), is completely regular). The uniformity is uniquely determined, if we require it to have a base of translation-invariant vicinities. Proof. We call a vicinity N<^ExE translation-invariant if (x + z, y + z) lies in N whenever (x,y) does, for arbitrary z in E. The uniformity on E[X] is defined by the vicinities Nv consisting of all pairs (x,y)eExE with y — xeU, UeU. We show that the Nv form the base of a Hausdorff uniformity. The Nv form a filter base on ExE which clearly satisfies (V1) of § 5,1., and also satisfies (V4) of § 5, No. 2, since f]U = o. For circled U, U= — U, so that Nv is symmetric, i.e. Nu = N{j'1, and (V27) is satisfied. Finally it follows from the existence of a circled V with V+ V^U that Ny <= Nv, so that (V3) is satisfied. The topology £ is evidently the topology defined by this uniform structure. Finally, suppose that some translation-invariant uniformity is given on E for which £ is the corresponding topology. If U is a neighbourhood of o which is obtained from a translation-invariant vicinity N, so that U is the set of all y with (o,y)eN, then x+U is the set of all z with (x,z)eN, i.e. N is equal to the set Nv defined above, and the uniformity is uniquely determined. 10*
148 § 15. Topological vector spaces Theorem (3) allows us to apply the results of § 5 about uniform spaces to topological vector spaces. (4) If a linear mapping A of E[jXi] into £[jX2] f5 continuous at o, it is continuous everywhere, and indeed is uniformly continuous. For the proof, cf. §10,2.(9). (5) The mapping (x,y)^>x + y from £[JI]x£[jX] into £[£] is uniformly continuous. Proof. Let V+V<=U, and let Nv be the vicinity of all (x,xr) with x' — xeV. Then the image of the vicinity NvxNv lies in Nv, since the differences (x'+/)-(x+j) of the (x,x\y,y')eNv x Nv lies in U. Trivially, (6) Every linear subspace H of a topological vector space is again a topological vector space, under the induced topology. (7) The closure H of a linear subspace H is again a linear subspace. Let x0 and y0 be closure points of H and let U be a neighbourhood of o. There exists V+V<=U, and so if x, yeH and xex0+V, yey0 + V, then x + yex0 + y0 + U. Hence x0+y0 is a closure point of the x + yeH. Further if x0 is a closure point of the points xeH, Xx0 is a closure point of the points Xx. 3. The completion. Every normed space can be completed to give a (B)-space (§14,3.(1)); in the same way the following result holds for arbitrary topological vector spaces: (1) Every topological vector space £[£] over K can be embedded in a smallest complete topological vector space £[£], the completion of £[£]. This is unique up to topological isomorphism. The closures in E[i] of the neighbourhoods of a base of neighbourhoods of o in E[%~\ form a base of neighbourhoods of o in £[£]. The closure of a circled neighbourhood is again circled. Proof. £[£] is a uniform space, by 2. (3). By § 5, 5.(2), there is a smallest completion EpX], uniquely determined up to uniform space isomorphism, whose points consist of equivalence classes of Cauchy nets. E[%~] is again Hausdorff. By 2.(5), addition x + y is a uniformly continuous mapping from E\X\ x E[X] into £[£]; it is also uniformly continuous as a mapping from E[2]xE[a:] into £[2]. By §5,4.(4) and §5,7. it can be extended uniquely to a uniformly continuous mapping from £[£] x E[pL\. By 2.(4) ax is also uniformly continuous on £[£], for fixed a; scalar multiplication can therefore also be^ extended uniquely, in a uniformly continuous way, to the whole of £[£]. Axioms (LI) and (L2) for a vector space now follow for £ by taking limits in the axioms for £.
4. Quotient spaces and topological products 149 If {Nv} is the vicinity basis derived from a base {U} of neighbourhoods of o of EpX], then by §5,5.(4) the closures Nv in ExE form a base of vicinities in £[£]. We shall show_that Nv = Nv, where U is the closure of U in E[pL~\. A point (x,y)eNv is a closure point of pairs (x,y)eNv. But it follows from y — xeU that y — xeU, so that NvczNv. Conversely if (x,y)eNfj, so that z = y — xeU, then z is a closure point of points zeU and 3c is a closure point of points xeE; hence (x,y) = (3c,3c + z)_ is a closure point of elements (x,x + z)eNv, implying that NfjCiNo. The closures U in £[£] of the neighbourhoods U therefore form a base of neighbourhoods ofo in £[£]. If the sets U are circled, so also are the sets U9 for from <xU a U(\a\^l) it follows by continuity that olU a U. From V+ V<=. U and the continuity of (x,y)^>x + y it follows that V+VaU. Finally every U is absorbent, as well: given x0eE there exists xeE with x0 — xeV; if xenV, then x0 = (x0 — x) + xe V+n K<= nU. Applying 2.(2), E[i~\ is therefore a topological vector space. Two smallest completions of E[X~\ are isomorphic as uniform spaces; because of the continuity of the vector space operations, sums and scalar multiples correspond, so that the uniform space isomorphism is a topological isomorphism of topological vector spaces. This completes the proof of (1). Remark. The completion of a linearly topologized space (cf. § 13,3.) can be constructed in exactly the same way. In this case U is a linear subspace; we conclude from this that U is also a linear subspace and, using § 10,2.(1), obtain the result corresponding to (1). 4. Quotient spaces and topological products. As for normed spaces and linearly topologized spaces we define the induced topology lona quotient space E/H of a topological vector space E[X] by taking the images K(0) of the open sets 0 of E as open sets (where K is the canonical mapping of E onto E/H). (1) // H is a closed linear subspace of the topology vector space EpX], E/H is a topological vector space under the induced topology. A similar argument to that of § 10, 7. shows that the sets K(0) determine a Hausdorff topology on E/H. If U c E is absorbent and circled, K(U) is also absorbent and circled, and it follows from K+Kci U that K(V) + K(V) a K(U). If, therefore, we start from a base of neighbourhoods of o in E which satisfies the requirements of 2.(2), we obtain a base for the induced topology on E/H with the same properties, using the canonical mapping. The assertion now follows from 2.(2). We remark that E/H need not be complete if E is. For counterexamples see § 23, 5. and § 31, 6.
150 § 15. Topological vector spaces The concepts of topological homomorphism, topological monomorphism and topological isomorphism are defined as in § 10, 7., and the proofs given there establish the three following results: (2) The canonical mapping K of £[JX] onto E/H, with H closed, is a topological homomorphism. (3) A continuous linear mapping A of £[JXi] into FpX2] always has a closed null space N\_A~\. A is the product of the canonical homomorphism K of E onto E/N\_A~\, a continuous one-one linear mapping A of E/N\_A~] onto A(E) and the embedding J of A(E) into F. (4) Every topological homomorphism of £[JXi] into F[jX2] *5 the product of the canonical homomorphism K of E onto E/N\_A\ and a topological monomorphism A of E/N\_A~\ into F. The topological monomorphism A is the product of a topological isomorphism A of E/N\_A~\ onto A(E\ and the embedding J of A(E) into F, which is a topological monomorphism. Just as in § 10,7. if F a G are two closed linear subspaces of E[X] we have a topological isomorphism (E/F)/(G/F) ^ E/G. In the converse direction we have (5) A linear mapping A of £[^i] into F[U2] with a closed nullspace is continuous (respectively a topological homomorphism) if A is continuous (a topological isomorphism). In § 14,4. we could only show that the topological product of finitely many normed spaces is again normable. Now we have, more generally, (6) The topological product E [%] = TT Ea pXJ of arbitrarily many a. topological vector spaces £a[jXa] is again a topological vector space. To prove this, we use §1,8. and § 7, 8. It is not difficult to verify that the conditions of 2.(2) are satisfied by the base of neighbourhoods U = TT Wa, where finitely many Wa are circled neighbourhoods Ua of o in £a, and all the other W« = Ea. (7) The topological product of infinitely many normed spaces is not normable. For a neighbourhood ||x||<l must certainly contain a neighbourhood TT Wa. Now let x be an element all of whose components xa = o, a except for an xp which lies in a Wp which is equal to Ep. But then all scalar multiples of x must also lie in TT Wa\ since ||/lx|| = \X\ \\x\\ < 1 for a all AeK, we obtain ||x|| = 0, which is a contradiction. We have already shown, in § 5, 7., that the completion of a topological product is equal to the topological product of the completions, so that E[X] = TT £a[jXa] is complete if the £a[jXa] are.
5. Finite dimensional topological vector spaces 151 As in § 7, 8., we denote the topological product of d copies of the field K by cod(K). If d = K0, we simply write co(K). Since we no longer take the discrete topology on the field of real or complex numbers, but take the natural topology instead, we obtain a topology on cod which is different from the linear topology. By (7), od is only normable if d is finite. 5. Finite dimensional topological vector spaces. The ^-dimensional real or complex space K" is a topological vector space under its natural topology. The fact that this is the only topology on K" which is compatible with the vector space operations was established by Tychonoff [ 1 ]. Thus we have (1) Every n-dimensional topological vector space E[X] over K is topologically isomorphic to K" with its natural topology. Proof, a) We can identity the vector space E with K". Let U be a circled ^-neighbourhood of o, and let V be a circled ^-neighbourhood of o for which the n-fold sum V+ ••• +V is contained in U. Since V is absorbent, there is a k>0 such that all the feef, i=l,...,n, lie in V9 n where the e, are the unit vectors. But then all the terms k £ a,-ef with i= 1 Y, M2^ 1 lie in U, i.e. U contains the Euclidean ball of radius k determined by taking the ef as unit vectors. The topology £ is therefore coarser than the natural topology on K". b) In order to show that £ is finer than the natural topology on K" it is enough to show that there is a ^-neighbourhood of o which is bounded in K", for then it lies in some Euclidean ball. Let U0^fE be a circled neighbourhood of o. U0 can contain a linear subspace of dimension at most n—\. If V is a circled neighbourhood of o with K+Kci U0, V can only contain one linear subspace //„_! of dimension n— 1, for otherwise U0 would be equal to E. We now take a circled neighbourhood W of o which does not contain some non-zero x in Hn_1. Then U1 = VnW can contain linear subspaces of dimension at most n — 2. Continuing this process, we obtain a circled neighbourhood U of o which is bounded along every straight line through o. Because E[X] is regular, we can suppose that U is closed. If U were not bounded in K", we could find a sequence x{p) of vectors with Euclidean lengths ||x(p)|| = l and with x(p) e— U. Some subsequence of P this would then converge in K" to a vector x(0)#=o, and by a) it would also converge in the topology X. Because U is closed, x{0) would lie in 1 ^ 1 each — U, and would therefore lie in \)—U. But this intersection is o, P p P
152 § 15. Topological vector spaces since U is bounded on every straight line through o. This contradicts the fact that x(0)=ho. As a immediate consequence of the fact that K" is complete we have (2) Every finite-dimensional linear subspace of a topological vector space is closed. As in § 10,4.(7), we have (3) // F is a closed linear subspace and G a finite dimensional linear subspace of a topological vector space E[X], then F + G is closed. We cannot apply the theory of duality to prove this result, as we did in § 10. Let K be the canonical mapping of E onto E/F. The image K(G) is finite-dimensional in E/F, and so is closed, by (2). Because K is continuous, the inverse image Ki~1)(K(G)) = F + G is closed. We recall that every linearly topologized space is totally disconnected. In contrast, we have (4) Every topological vector space E[X] over K is connected. By (1), every straight line through o in £[£] is topologically isomorphic to K. The straight line joining x0 and y0 results from a parallel displacement of the line through o and y0 — x0; it is also topologically isomorphic to K, by 1.(1), and it is therefore connected. The connectedness of E[X~\ follows from this, using §1,6. 6. Bounded and compact subsets. A subset B of a topological vector space £[£] is said to be bounded (or ^-bounded, if we wish to specify the topology) if for each neighbourhood U of o there is a p>0 with BapU. In a normed space, this clearly means sup||x||<oo, in agreement with the definition of § 4, No. 2. xeB If x0 is a non-zero element of E[X~] and if U is a circled neighbourhood of o which does not contain x0, then the only scalar multiples ax0 which can lie in U are those with |a| < 1. If B is bounded and£ c pU, then the only scalar multiples ax0 which can lie in B are those with |a| <p\ a bounded set therefore cuts a straight line through o in a subset of a finite interval. Every subset of a bounded set is bounded. Every finite set is bounded. The circled cover of a bounded set is bounded. (1) The sum and the union of finitely many bounded sets are bounded. We show this for two bounded sets B1 and B2. Let U be circled; if Bxa p1U,B2<= p2U, and if p = max(p1,p2) then Bxv B2<^ pU. Further let K+Kc[/, and as before, let B^ pV,B2<=. pV. Then Bl+B2apV + pVapU.
6. Bounded and compact subsets 153 (2) The closure B of a bounded set B is again bounded. For BczpU follows from BczpU; but the sets U form a base of neighbourhoods ofo, since £[£] is regular; hence B is bounded. (3) A subset B a E[%] is bounded if and only if whenever xn is a sequence in B and a„-»0 in K then a„x„->o in £[£]. a) Let B be bounded and let U be a circled neighbourhood of o. There is a p>0 such that xnepU for all n. Then <xnxneocnp U = \ocn\pU. If n0 is chosen sufficiently large, so that |a„|p^l for n^n0, then ocnxn lies in U, for n^n0. Since this holds for each U, ocnxn converges to o. b) If B is unbounded, there is a sequence xneB and a circled neigh- 1 bourhood U of o with xn$nU. The sequence — xn lies outside U and does not converge to o. n (4) In a topological vector space, every Cauchy sequence xn is bounded. Given a circled [/, let V+Va U, with V circled as well. All the terms xn — xno lie in V, for n^n0, so that xnexno+V. There exists a pV,p^.l, which contains all the elements xl9...,xno. But then all the xk9 fc=l,2,..., lie in pV+Vcz pU. (5) T/ze continuous linear image of a bounded set is bounded. Suppose that A is a continuous linear mapping of EpJ into F[jX2]- Given a neighbourhood U of o in F, there is a neighbourhood F of o in E such that A(V)cU. It then follows from BapV that X(B)c:pX(K)c:pC/. We now establish some results concerning precompact and compact subsets of a topological vector space £[£]. (6) Every precompact subset K of E[X] is bounded; so therefore is every compact subset. By §5, 6.(2), K is totally bounded, so that given a neighbourhood V ofo there are finitely many sets x{+V, xteK, which cover K. If p>\ n is chosen so that all the xf lie in p V, then we have K c (J (x. + K) cz pV+VapU, if K and U are circled and V+VaU. i = i For compact sets, the following result is contained in § 3,2.(5): (7) T/ze continuous linear image of a precompact set K is precompact. Let K be precompact in £[3^], and let A be a continuous linear mapping from £[3^] into F[jX2]- Given a neighbourhood U of o in FpX2], there is a neighbourhood V of o in E with A(V) c C/. If X c: (J (x,+ F), then X(K) c (J (^xf + ^(K)). It follows from this that i = 1 r = 1 ,4(K) is covered by n sets which are small of order Nv, so that A(K) is precompact.
154 § 15. Topological vector spaces (8) // Kl,...,Kn are precompact (respectively compact) subsets of E[X] and if a1?..., a„ are arbitrary constants in K, then the set ai Kx + ••• +otnKn is also precompact (compact). First we show this for compact Kt. Since a,/^, being the continuous image of Kh is again compact, it is sufficient to prove the result with af=l. By Tychonoff's theorem the topological product Kx x • • • x Kn is n compact. The mapping (xt,..., xn) -> £ xt from K1 x • • • x Kn into i= i E[X~\ is continuous, and so the image K1 + ••• + Kn is compact. If theJC, are precompact, we consider the completion E[X~\. The closures Kt in E[X~\ are compact, and, as we have just shown, so is alKl + ••• +<xnKn = ((x1K1 + ••• +anKn); i.e. cctKx + ••■ +oinKn is precompact. The union of finitely many compact (respectively precompact) sets is compact (precompact). (9) // K is compact and disjoint from the closed subset M of EpX], there is a neighbourhood U of o for which (K+U)n (M + U) is also empty. Proof. For each yeK there is an open circled neighbourhood Vy of o for which {y+Vy+Vy+Vy)nM is empty, (y + Vy + Vy) n (M + Vy) is also empty. K is covered by finitely many yf+Ky.. Let U = f]Vy.. Then for each yeK we have ' y+Uczy^Vn+Uczyi+Vn+V^ so that (y+ U)n(M + U) is empty, for all yeK. (10) // K is compact and M is closed in £[£], then K + M is closed. If x$K + M, (x-M)nK is empty. But x — M is also closed, so that by (9) there is a U with ((x-M)+U)nK empty. Hence (x + U) n(K + M) is empty, and so the complement of K + M is open. (11) If K is closed, precompact and disjoint from the complete subset M of £[£], then there is a neighbourhood U of o for which (K+U) n(M + U) is also empty. The closure K of K in the completion E[i~\ is compact, and is still disjoint from M. (9) proves the result for the subsets K and M of E; the result for the subsets K and M of E follows from this. In the same way, it follows from (10) that (12) // K is precompact and closed, and if M is complete in £[£], then K + M is closed. As an example of the concept of boundedness, we consider topological products.
8. Topologically complementary spaces 155 (13) The bounded sets of a topological product E[X~\ = T\ Ea[Xa~] a are just the subsets of sets of the form T7J3a, with Ba bounded in £a[£J. a If B is bounded in £, then by (5) the projections Ba of £ are ^-bounded in £a[Ia], and so B c T7£a. Conversely a set T\Ba is bounded in £[£]; a a. this follows directly from the definition of the product topology. Since every bounded set in the field K is relatively compact, it follows from Tychonoff's theorem that every bounded set in cod(K) is relatively compact. A set B of vectors x = (£a) in a>d is bounded if it is "coordinate- wise" bounded, i.e. if for each a there is an Ma>0 for which \£a\^Ma for all xeB. 7. Locally compact topological vector spaces. In 5. we showed that every finite-dimensional topological vector space is topologically isomorphic to K", and so is locally compact. The following converse holds: (1) Every locally precompact topological vector space is finite dimensional. The hypothesis means that E[X] has an open precompact circled neighbourhood U of o. Starting with H0=o, we construct a strictly increasing sequence HQ(^Hlcz ••• of finite-dimensional linear sub- spaces of E. Suppose that Hk has already been constructed, and that it is a proper subspace of E. Hk is closed by 5.(2). By 6.(6), U is bounded, and it is mapped into a bounded subset of E/Hk by the canonical mapping K of E onto E/Hk9 by 6.(5). K(U), being bounded, is a proper subset of E/Hk, so that its inverse image K(~l)(K(U)) = Hk+ U is a proper subset of E. We now show that the closure U of U is not contained in Hk + U. If C7 c= Hk +JJ9 then we would have Hk+U = Hk + U. But by (5.(1) and 6.(12), Hk+ U is closed. On the other hand, Hk+ U is open, so that E would possess a proper open and closed subset, contradicting 5.(4). There therefore exists a point ykeU which does not lie in Hk+U. We now put Hk + 1=Hk® \yk~\. If the sequence of spaces Hk constructed in this way were not to produce E after finitely many steps, we would obtain a sequence ykeU with y^ — y^U for z=f=/c. There exists a circled V with F + Kc[/. Since U is totally bounded, it is covered by finitely many x,+ K But then at least two yh yh and yh say, would lie in the same Xj+V, and so yh— yi2eV+ V a U, which is impossible. 8. Topologically complementary spaces. We now link up with the ideas of §10,7. Two closed algebraically complementary linear sub- spaces Hx and H2 of the topological vector space E[X~\ are said to
156 § 15. Topological vector spaces be topologically complementary if the mapping (xl,x2)->xl +x2 of the topological product HX[X~\ x H2[X~\ onto E[X~\ is a homeo- morphism. We repeat a result which was also valid before (§ 10,7.(6)): (1) A closed linear subspace Hx of E[X~\ has a topological complement H2 if and only if there is a continuous projection Px of E onto Hv Px is then a topological homomorphism, H2 = N\_P1], and E/H2 is topologically isomorphic to Hv The proof is almost word for word the same as in § 10, 7., and is left to the reader. (2) A closed linear subspace H of finite codimension in a topological vector space £[£] always has a topological complement. Every algebraic complement of H is also a topological complement. Proof. Let M be an algebraic complement of//, so that E = H ©M. By 5.(2), M is closed and is topologically isomorphic to K". The projection P from E onto M with null-space H is the product of the canonical homomorphism K of E onto E/H, a one-one linear mapping P of E/H onto M and the embedding J of M into E. But since E/H is also topologically isomorphic to K", P is a topological isomorphism, and so, by 4.(5), P = JPK is continuous, and the assertion follows from (1). The obvious conjecture (cf. § 10,7.(8)) that every finite-dimensional linear subspace of E [X~\ always has a topological complement is false, as we shall see in the next number. 9. The dual space, hyperplanes, the spaces LP with 0<p<l. We again denote the vector space of all continuous linear functionals on E by E', and E is again called the dual of E. In this number we shall show, by giving an example, that there are topological vector spaces on which there are no continuous linear functionals other than the trivial one u(x) = 0, so that the dual space can consists of the zero element alone. This pathological possibility prevents us from establishing a really meaningful and useful theory of general topological vector spaces. Such a theory is only obtained when we restrict attention to locally convex spaces, as we shall do in the next chapter. The following important relation holds between continuous linear functionals and closed hyperplanes. (1) // u(x) is a linear functional on E[X~\ which is not identically zero, its null-space is a hyperplane in E; conversely, for each hyperplane through o in E there is a linear functional whose null-space is the hyperplane.
9. The dual space, hyperplanes, the spaces LP with 0<p< 1 157 u(x) is continuous if and only if the corresponding hyper plane is closed. A (closed) hyperplane in E[X~\ is given by an equation (2) u(x) = y, yeK, where u(x) is a (continuous) linear functional. A hyperplane which is not closed is dense in E[X~\. Proof, a) If w=f=o is a linear functional on £, H = u~1(o) is a vector space, which is closed if u is continuous. H has codimension 1 in E: there is an x0eE with u(x0)=l; for arbitrary x in E we have a unique decomposition (3) x = u(x) x0 + (x - u(x) x0), where the second summand lies in H\ the scalar multiples of x0 therefore form a one-dimensional complement to H. b) Let H be a hyperplane through o. Let E = \_x0~\@H. Let us put w(ax0 + y) = a for yeH. Then u is a linear functional on E with null- space H. If H is closed u is continuous, for the mapping it from E/H onto K is the topological isomorphism ax0-»a (cf. 4.(5)). c) A general hyperplane is of the form x0 + H, where H is a hyperplane through o; it is closed, if and only if H is. If u is a characteristic linear functional for H, then u(x0 + y) = u(x0) = y holds for all yeH and for no other elements of E. d) If H is not closed, H is a proper subspace of H. Since // has codimension 1 in E, H must equal £. The following example was, originally given by M. Day [1]; the simple proof of (9) was given by W. Robertson [1]. Once again let LP be the space of all measurable functions f(t) on I = [a,b~\ with J I/WIP^<°° (again, we identify functions which are equal almost everywhere); a the case / = (— oo, oo) is also valid in what follows. This time, however, let 0 < p < 1. A topology is determined on LP by taking as base of neighbourhoods of o the sets U{c) of all fell with (4) ll/llP = (fl/(Olp^Y/P^fi. Clearly ||a/||p=|a| ||/||p. The sets U(e) are circled and absorbent, and form a filter-base on LP with f]U(£)=o. LP is a topological vector space, if we can show that for each U there is a V with V+ Kc= U. To this end we show that instead of Minkowski's inequality we have the following inequality: (5) \\f+g\\P^2~^(\\f\\p+\\g\\p).
158 § 15. Topological vector spaces Proof. For q>\ the function has exactly one minimum for x>0, namely where x=l, so that 1 + xq^2l~q(l+x)q. Putting x = c/d, it follows from this that if c, d> 0, then (6) cq + dq^21-q(c + d)q (q>\). But it follows from (c + d)p^cp + dp, for 0<p<\, and from (6), with q = \/p, that x x Thus if U is one of the neighbourhoods (4), it follows from (5) that (7) 2-llpU + 2-l/pU^U; hence (8) LP, 0<p< 1, is a topological vector space. Now let u(f) be a continuous linear functional on LP which does not vanish identically. There exists g0eLp with u(g0)=\. For a<s<b we put ^(O equal to the function which is equal to g0(t) in the interval a^t^s, and which vanishes s identically for t>s; let g{s2)(t) = g0(t)-g{s2){t). Now \\g(sl)\\p = i\ao{t)\pdt increases continuously from 0 to \\g0\\p, so that there is an s0 with H^II^II^H^illfifoli;. Since \u(g0)\ = l, we must have MfiOl^i for i=\ or 2. Let gl{t) = 2g(s,)0{t), for this i. Then ^(gfjl^l, but ||gf1||p = 2 p||fif0llp- Repeating this procedure we obtain a sequence g„ with \u(gn)\^l, but with ||#Jp = 2 p \\g0\\p->09 contradicting the continuity of u. Thus we have shown: (9) Every continuous linear functional on LP, 0</?<l, vanishes identically. Using this, we can answer the problem raised in the preceding number. (10) In LP, 0<p< 1, every hyper plane is dense, and no finite-dimensional linear subspace H has a topological complement in LP. Let Hi be a one-dimensional linear subspace of H. By 5.(1), Hi has a topological complement H2 in H under the topology induced by LP, so that H = Hl@H2. If G were a topological complement of H, H2®G would be a topological complement of Hl; H2®G would therefore be a closed hyperplane of LP, which is not possible, by (1) and (9). We establish one further general result about dual spaces: (11) A topological vector space E[X~\ and its completion E\_T~\ have the same dual space, so that E' = {E)'. Since E is dense in E, and since every continuous linear functional u on E is uniformly continuous, by 2.(4), u can be extended in a unique way to a uniformly continuous functional on E\ conversely the restriction to E of a continuous linear functional defined on E is also continuous on E.
10. Locally bounded spaces, quasi-norms, p-norms 159 10. Locally bounded spaces, quasi-norms, p-norms. The expression (4) of the preceding number does not possess all the properties of a norm. It is an example of a quasi-norm ||x|| on a topological vector space £[£], which is characterized by the following properties: (Qi) IMI^O, (Q2) x=o if ||x||=o, (Q3) ||ax|| = |a|||x|| for all aeK, (Q4) There is a fc^l for which ||x + j;||^/c(||x|| + ||j;||) for all x,ye£[3;]. If it possible to take k= 1, we obtain a norm. A base of neighbourhoods of o in Lp, 0<p<l, is given by the sets ||x||p^£. Those spaces whose topology can be given by a quasi-norm in this way can be characterized quite simply. A topological vector space is said to be locally bounded if there is a bounded neighbourhood of o. This can be taken to be circled. We have (Hyers [1], Bourgin [1]): (1) The topology of a topological vector space E[X] can be given by a quasi-norm if E[X~\ is locally bounded. Conversely, a quasi-normed space is always locally bounded. Proof, a) Let U be a circled bounded neighbourhood of o in £[3f|. The sets olU, a>0, form a base of neighbourhoods of o, for U is bounded, so that for each neighbourhood V of o there exists /?>0 for which 01/ <= K We introduce the so-called Minkowski functional q(x) of U: (2) q(x)= inf a (a^O). xectU q(x) is well-defined for each x, for since U is absorbent there always exists an a>0 with xeolU. We now show that q(x) has the properties of a quasi-norm. (Q 1) is trivially satisfied, and (Q2) follows from the fact that a non-zero x cannot lie in every at/, since f] aU=o. (Q3) follows from the fact that U is circled. a>0 Since the sets a (7, a>0, form a base of neighbourhoods, there exists k>0 with U + UczkU, by (LT1). Now let q(x) = p09 q{y) = a0. Then if p>p0 and o>oQ the elements x/p and yjo lie in U, so that (3) -J^* + ^y = ?±leU+U. p + a p p + a o p + G It follows from this that x + yek(p + (j)U, so that q(x + y)^k(q(x) + q(yj); we have therefore established (Q4).
160 § 15. Topological vector spaces If Kis the set of all xeE with g(x):gl, then we have U cz Fc(l +s)U, and so the quasi-norm topology coincides with X. b) If a quasi-norm ||x|| is defined on E, and if we introduce a topology X' on E by means of the absorbent circled sets a V, a > 0, consisting of all x with ||x||^a, then we obtain a topological vector space, since f] aV=o and since — V + — V c K, by (Q4). Finally, V is a>0 ^ ^ clearly i'-bounded. In the next paragraph we shall investigate the concept of convex set in detail. For locally bounded spaces, a generalization of this concept, introduced by M. Landsberg, has proved to be important. Let 0<p^ 1. A subset M of a vector space E over K is said to be p-convex if whenever it contains x and y it contains xx + oy, where t^O, 0-^0 and xPjrop=\. M is said to be absolutely p-convex if whenever it contains x and y it contains all xx + oy with |t|p + |o"|p^1, where t and a are real or complex, depending on whether E is a real or complex vector space. If p=l, we speak of convex and absolutely convex sets, respectively. The absolutely p-convex cover \~P(M) of a set M is the intersection of all the absolutely p-convex sets which contain M. \~P(M) n n consists of all the terms of the form £ afx,, with ]T |af|p^l. The i=l i=l proof for p with 0<p<l is completely similar to the one for p=\ given in § 16,1.(1), using the inequality |a + jft|p^|a|p + |jft|p. A p-norm |||x|||, 0<p:gl, on a vector space E is characterized by (Pl) IIMH^O, (P2) if |||;e||| = 0, then x=o, (P3) |||ax||| = |aH||x|||, for all aeK, (P4) |||x + ),|||^|||x||| + ||M||. If p = 1, we obtain the concept of a norm. The p-th power of the quasi-norm ||/||p in LP (cf. 9.(4)) is a p-norm HI/HI on Lp, and the neighbourhood 9.(4) of o is p-convex. The topology on LP can also be given by the neighbourhoods |||/|||^c; we say that LP is a p-normable topological vector space. We now establish (4) A topological vector space E[X~\ is p-normable, 0<p^l, if and only if it has a p-convex bounded neighbourhood of o. In the case p=l, this is the characterization of normed spaces given by A. Kolmogoroff [1].
10. Locally bounded spaces, quasi-norms, p-norms 161 Proof. It follows directly from (PI) to (P4) that in a p-normed space the set U of all x with |||x|||^ 1 is absolutely p-convex; further U is bounded, since WpV^III^P holds for p>0, by (P3), so that pl/pU lies in the neighbourhood |||x|||^p of o. Conversely, let V be a bounded p-convex neighbourhood of o in E[X~\. V contains a circled neighbourhood W of o. Let \~P(W) be the n absolutely p-convex cover of W, and let ]T af xf, with x^e W, £|af|p^ 1, i=l be an element of \~P{W). We put ]T|af|p = pp. The element ^afx,- = Y —— fifXf, with ef = —-p, is of the form V— — v,-, j/fe Wc: K, since W P |af| _ p ' is circled. Now if a p-convex set contains z1,...,zM, it always contains n n £ (j.Zi, where /?^0 and £ j8f=l (this is proved analogously to i = 1 i = 1 " la-l § 16,1.(1)). Since the yt lie in K, Y, ~yt a^so ^es m ^ This means that i = l P rp(W0 <= K We can therefore assume that V is absolutely p-convex. We put |||x||| = inf pp, p^O; this is equal to q(x)p, where #(x) is xepV the quasi-norm of V. (PI) to (P3) therefore follow from (Q 1) to (Q3). If x and y are arbitrary points in £, and if p>p0 = q(x), a>a0 = q(y), then because V is absolutely p-convex the element p x ay x + y (pp + <rp)l/p'~p + (pp + ap)l/p ~o = (pp + ap)l/p lies in V, and (P4) follows easily from this; |||x||| is therefore a p-norm, and it follows from V a U <= (1 +c)V, where U is the set |||x|||^l, that X is defined by the p-norm. Every p-normed space is locally bounded. We shall show that conversely every locally bounded space E[X~\ is p-normable for a suitable p with 0<p^l (cf. S. Rolewicz [1]). For every bounded circled neighbourhood U of o in E[X], there exists at least one /c^2 with U +U akU. The greatest lower bound of these k is called the module of concavity of U. The greatest lower bound of all these modules of concavity is called the module of concavity k(E) of E[X~\. We now have (5) // k(E) = 21/P0 is the module of concavity of the locally bounded space E[_X~], then for each p<p0 there is a p-norm which determines the topology X. 11 Kothe, Topological Vector Spaces I
162 § 15. Topological vector spaces By (4), it is sufficient to show that £[£] has a bounded absolutely p-convex neighbourhood of o. By hypothesis there is a bounded circled neighbourhood U of o with 1 _ 1 _ A U+U<=2PU, i.e. 2 pU + 2 PU<=U. More generally we have _ *I _ hn n (6) 2 p£/ + "- + 2 pU^U{ox Y,2~ki^l> where the/c,. are positive integers. i It is sufficient to prove this for £2"k,"=l. By the order of such a decomposition of 1 we mean the maximum of the kt which appear. The assertion is correct for decompositions of order 1. Every decomposition of order k +1 results from one of order k by replacing certain summands 2~k by 2_(fe+1) + 2~(fe+1), since it follows from ^2~fe,= l that an even number of summands 2~{k+1) appear. Making the inductive hypothesis that the relation (6) is true for k, we obtain the relation for _ k fc+1 when we replace the corresponding summands 2 p U by _k+1 _k+1 _ 1 _ 1 2 p U + 2 p U; this is permissible since 2 pU + 2 PU<=U. 1 « We now show that rP(U) <= 2P U, which establishes (5). Let £ |af|p^l, != 1 and let kt be chosen so that 2~kl^\(xi\p<2~ki+l holds. But then n n _1 _ki J_ YJMp<2YJ2-kl^2, so that Juaixie2PYJ2 pUc2pU for arbitrary i i xteU, by (6). It can be shown, by giving examples, that E\_X~\ need not be p0- normable; Lp, 0<p^l, is /?-normable, however, and k(LP) = 21/p (cf. Rolewicz [1]). (7) // E[X~\ is locally bounded, then so is every closed subspace and every quotient space, and so also is the completion E[X"]. We shall only prove the last assertion. If U is a bounded neighbourhood of o, the sets p U, p > 0, form a base of neighbourhoods of o in E. By 3.p) the closures p U in E[Z~\ form a base of neighbourhoods of o in £[£]; i.e. U is a bounded neighbourhood ofo in E[X~\. Examples and further results about locally bounded spaces are contained in the papers of Bourgin [1], Hyers [1], Landsberg [1], [2] and Rolewicz [1]. 11. Metrizable spaces. In the classical theory of Banach, not only were normed spaces considered, but also those spaces on which a metric is defined which is compatible with the vector space operations. First we consider the question of when the topology of a topological vector
11. Metrizable spaces 163 space £[£] can be given by a metric. In this case we say that E[X~\ is metrizable. (1) A topological vector space is metrizable if and only if it has a countable base of neighbourhoods of o, 50 that it satisfies the first axiom of countability. If this is the case, there is always a translation-invariant metric, and this defines the uniform structure of £[£] as well. E[X~\ is only metrizable if there exists a countable base of neighbourhoods of o. If this is the case, then, by §6,7.(1), the translation- invariant uniform structure of £[£] (and not just the topology) is metrizable. We now show that the metric constructed in §6,7.(2) is translation-invariant. A base of vicinities of the uniformity of £[£] is given by the sets Nv, as U runs through the circled neighbourhoods of o. Following through the proof of §6, 7.(2) once again, we see that, since (x,y) and (x — y,o) always belong to the same vicinities, the function f(x,y) defined there satisfies f(x,y) =/(x —y,o); further, if we make the sequences x = xl5 x2,..-,x„ = y and x — y, x2 — y,...,y — y correspond to each other, it follows that |x,y| = |x — y,o|. The metric is therefore translation-invariant. In addition let us remark that, as each U is circled, ax belongs to U for each \a\^1 if x does; hence /(lx,o)^/(x,o), so that |/x,o| ^|x,o|. In particular it follows from this that |tx,o|=|x,o| if |t| = 1. Once again we write ||x|| = |x,o|, and obtain (2) The uniform structure of a metrizable topological vector space can be given by a function ||x|| with the following properties: (Fl) llxll^O, (F2) x=o if ||x|| = 0, (F3) Ux\\£\\x\\ if \A\^h (F4) Hx + j^llxH + llyll, (F5) ||/x„H0 if ||xJ-0, (F6) IkxHO if /B->0. A function ||x|| with properties (Fl) to (F6) is called an (F)-norm, and E is said to be (F)-normed. The sets Vr of all x with ||x|| <e form a base of neighbourhoods of o for the topology determined by the (F)-norm. (F5) and (F6) result from the fact that /.xn-*o and a„x->o in the topology % so that |/x„,o|->0 and |/„x,o|-*0. The converse of (2) is also valid: (3) Under the topology defined by the (F)-norm, an (F)-normed space E is a topological vector space with a countable base of neighbourhoods of o. 11*
164 § 15. Topological vector spaces For by (F3) the sets VE are circled, and by (F6) they are absorbent. These sets clearly form a filter-base, whose intersection is o, because of (F2). Further, by (F4), VE/2 + Vrj2 cz VE. The sets V1/n form a countable base of neighbourhoods ofo. We remark that the uniform continuity of ||x|| on E can be established as in § 14.1, and that the (F)-norm of E can be extended continuously to the completion of E. (Fl) to (F6) still hold, as is shown by taking limits. If we define an (F)-norm on a quotient space E/H by setting ||x|| = inf ||x||, as in § 14,4., then it is easy to confirm that (Fl) to (F6) xex still hold for ||x|| provided that H is closed, and that the topology defined by this (F)-norm is the quotient-space topology. (4) The quotient space E/H of a complete metrizable space by a closed subspace H is again complete. The proof of § 14,4.(3) also holds for this more general case. A subset M of an (F)-normed space is bounded (in the sense of 6.) only if sup ||x|| < oo. As we shall see later, the converse is not true. Sets which are bounded in the metric sense (cf. § 4, 2.) need not therefore be bounded in the sense of 6. Banach called a complete (F)-normed space an (F)-space, after Frechet. Following the terminology of Bourbaki we shall reserve this name for complete metrizable locally convex spaces (cf. §18,2.). The p-norms on locally bounded spaces which were introduced in 10. are also (F)-norms, as can be seen immediately. We have introduced an (F)-norm ||x|| on every metrizable space £pl], and the corresponding metric defines the uniquely determined translation-invariant uniformity of E\_X~\ (cf. 2.(3)). Two different (F)-norms corresponding to the same topology on E\_X~\ are therefore always equivalent, i.e. the sets NE of all (x,y) with ||x — y\\<e form a base of vicinities for the uniform structure of £[2], for both (F)-norms. Conversely, instead of starting from a topology X on £, we can also start from a metric XR. If this metric is compatible with the vector space operations on £, i.e. if the mappings {x,y)->x + y and (a,x)->ax are continuous in the sense of the metric XR, then E[X~\ is called a linear metric space. XR defines a topology X on £, and E[X] is clearly a metrizable topological vector space. If XR is translation-invariant, then the uniform structures on E determined by SR and by X are the same, by 2.(3). If SR is not translation- invariant, this need not be the case. We consider the real line P as an example, first with the metric 5[R defined by the modulus (and so by an (F)-norm), and secondly with the metric 99^ defined by \a,fj\l = \<x3 — p3\. The vector space operations are compatible with 99^ as well, and Wl! defines the same topology on P as does 5[R. The uniform structure correspond-
11. Metrizable spaces 165 ing to 50i! with vicinities |a, /?11 < £, however, is strictly finer than the one corresponding to 50i, as can easily be seen. P is also complete with respect to 5tR1? since the 5[R-Cauchy and 90^ -Cauchy sequences in P are the same. A further metric on P is defined by |a,/?|2 = |tan-1a — tan_1/?|; this metric also gives the same topology, although P is no longer complete; e.g. the sequence oin = n is a Cauchy sequence which does not converge. The original metric of a linear metric space therefore need not give the uniform structure of the topological vector space; indeed the example given above is a case where E[50i] is not complete with respect to the metric 501, although it is complete as a topological vector space. Nevertheless, we have (5) If a linear metric space £[50^] is yjlrcomplete, then it is also complete as a topological vector space, and so is complete with respect to any of its (F)-norms. This conjecture of Banach was proved by V.L. Klee [3]. First we establish a lemma due to W. Sierpinski [1]: (6) Let H be a subset of a metric space £[501]. Let a second metric 50ix be defined on H, which induces the same topology on H as does 50i. //' H is 50i ^complete, then H is the intersection of countably many open subsets of £[50i]. Proof. Let xeH. Because 50i and 501 x induce the same topology on//, 1 1 there exists, for each n, a pn(x) < — for which \x,y\l < — holds for all yeH with \x,y\<pn(x). If Un(x) is the set of all zeE with |x,z|<p„(x), On= \J Un{x) xeH 00 is an open subset of £[50t] which contains H. If we put D= f] 0„, n= 1 then clearly H cz D. (6) is established, if we can show that D a H. Let z0eD. For each n there is an xneH with z0eUn(xn\ so that \z0,xn\<pn{xn) <—. Therefore z0 = limx„, in the sense of the metric n 901. We shall show that xn is a Cauchy sequence with respect to 9Jix. It follows from |xfc,x„|^|xfc, z0| + |z0,x„| < - + |z0,x„| and |z0,x„|<p„(xII) k j that |xk,xII|<pII(xII) for /c^/c0(n), and so |xfc,xw|1 < — for k^k0{n). 2 n Hence Ix^x^ fg— for k, /^/c0(n), so that xn is an 50irCauchy n sequence. Since H is (^Jll-complete, the sequence xn has a limit in H, and this must coincide with z0, so that z0eH and D cz H.
166 § 15. Topological vector spaces Proof of (5). £[501!] and the completion E of E with respect to the (F)-norm satisfy the hypotheses of (6). Thus £ is a dense linear sub- 00 space of E which is the intersection E = f] On of open subsets of E. 00 ~ n=l The complement E~E= [j (£~0„), being the union of countably «= i «, ~ many nowhere dense subsets, is meagre in E. If E~E were non-empty, it would contain a non-zero x0, and would contain x0 + £, so that E would also be meagre. This is impossible, since E = Ekj (E~E) is not meagre in itself (§ 4, 6. (5)). The assertion now follows from E = E. A detailed account of (F)-norms is to be found in Bessaga, Pelczynski and ROLEWICZ [1]. 12. The Banach-Schauder theorem and the closed-graph theorem. We now investigate when a continuous linear mapping from one complete metrizable topological vector space into another is a topological homomorphism. The fundamental Banach-Schauder theorem states that (1) A continuous linear m mapping A from one complete metrizable topological vector space E into another, F, either is a topological homomorphism or has an image A(E) which is meagre (of the first category) in A(E). (2) A is a topological homomorphism if and only if A(E) is closed. Proof of (1). It is sufficient to consider the case where A(E) = F, since A(E) is itself complete and metrizable. We assume that A(E) is not meagre. Let Up be the open ball of radius p about o in E, i. e. the set of all xeE with ||x|| <p, where ||x|| is an (F)-norm on E. First we show that the closure A(Up) contains a ball Va about o in F. 00 Since Up/2 is circled and absorbent, (J nUp/2 = E, and so A(E) = [j A(nUp/2). Since A(E) is not meagre, one of the sets A(nUp/2) n= 1 = nA(Up/2), and consequently A(Up/2) itself, is not nowhere dense in F. There therefore exists a yeA(Up/2), together with a neighbourhood V of o in F, for which y+V a A(Up/2). It follows from this that V cz -y + A(Up/2) cz A{Up/2) + A{Up/2) cz A{Up). However, V contains a ball V9. Next we show that if p'>p then A(UP) contains the ball Va. Suppose that y0eVa. We put p = pua = au and choose p2>P3>'" such 00 that YJPi<P' For each UPn there is a ball VGn in which A(UPn) is dense.
12. The Banach-Schauder theorem and the closed-graph theorem 167 We can take 0"„->O. Since A(UPl) is dense in V„l9 there is a xl with ll^iH <px for which the image A(xl) = yl satisfies \\y0— ^ill<o"2. A(UP2) is dense in Va2, and so there is an x2 with ||*2||<p2 f°r which the image A(x2) = y2 satis- II M II N II ^Pn + "- + Pm fies ||j;0 — yi— y2\\<^3, and so on. It follows from 00 and the completeness of E that the series ]T xt converges to an element oo i = 1 N x0; since £ Pi<p\ *o nes *n ^p'» anc* further /4(x0)= lim £>4(xf-) oo i= i N -*■ oc ^ = lim^<yI = >y0, so that yoe,4(t/p0. i We have therefore shown that the image of every open neighbourhood of o contains an open neighbourhood of o. This implies that the image of every open set is open; consequently A is a topological homo- morphism. Proof of (2). If A(E) is closed, A(E) is a complete metric space, and so by Baire's theorem (§4,6.(5)) it is not meagre in itself. A is therefore a topological homomorphism. If conversely A is a topological homomorphism, then by 4. (4) A determines a topological isomorphism A of the quotient space E/N\_A~\ onto A(E). But since E/N\_A~] is complete (11.(4)), A(E) is also complete, and so it is closed. By the graph G(A) of a linear mapping A from a topological vector space £[£i] into a topological vector space £[£2] we mean the linear subspace of the topological product £[^X1]xF[^I2] which consists of all pairs (x,Ax)9 xeE. The closed-graph theorem asserts that (3) A linear mapping A from a complete metrizable topological vector space E into a complete metrizable topological vector space F is continuous if and only if its graph G(A) is closed in Ex F. Proof, a) Suppose that A is continuous. A sequence (xn,Axn)eG(A) is a Cauchy sequence only if xn is a Cauchy sequence in E. Let x0 be the limit of xn. Because A is continuous, Ax0 is also the limit of the sequence Axn. Thus the Cauchy sequence (xn,Axn) has limit (x0,Ax0)9 and G(A) is closed. b) If conversely G(A) is closed, the projection (x,Ax)->x is a continuous one-one linear mapping from the complete metric space G{A) onto E, so that by (2) it is a topological isomorphism. It therefore follows from x„->o in £ that (xn9Axn)->o in ExF, and so Axn^o in F. A is therefore continuous at o, and consequently it is continuous everywhere.
168 § 15. Topological vector spaces The equivalent formulation given by Banach reads: (3') A linear mapping A from the complete metrizable space E into the complete metrizable space F is continuous if and only if whenever x„->x0 and Axn^>y0 then Ax0 = y0. The following assertion is more elementary: (4) A linear mapping A from a metrisable space E into a metrisable space F is continuous if x„->o always implies that Axn is bounded. Proof. It follows from (F4) that (5) ||fcx||^fc||x|| for each positive integer k. If xw->o, then ||xJ->0, and there is a sequence of positive integers /c„->oo for which fcII||xn||->0 as well. Applying (5), ||/c„xJ->0, and so the image sequence A(knxn) is bounded. Then — A(knxn) = A(x„)^>o, by 6.(3), which means that A is continuous at o. n We establish two corollaries of the Banach-Schauder theorem. (6) Two closed algebraically complementary linear subspaces Hx and H2 of a complete metrizable topological vector space E are topologically complementary. For the topological product Hx x H2 is again a complete metrizable space, and the mapping (x1,x2)->x1+x2 from HlxH2 onto E is one-one and continuous; it is therefore a topological isomorphism, by the Banach-Schauder theorem. (7) If a coarser metrizable topology X' is given on a complete metrizable vector space £[1], and if £[£'] is again complete, then 3/ is equal to 3. For the identity mapping of £[3] onto £[£'] is a topological isomorphism, by (2). 13. Equicontinuous mappings, and the theorems of Banach and Banach-Steinhaus. Let E be a topological space, F a uniform space, and 21 a collection of mappings A from E into F. The collection 21 is said to be equicontinuous at x0 if for each vicinity N in F there is a neighbourhood U{x0) in E such that (Ax,Ax0) lies in N for all xe U, Aetyi. 21 is said to be equicontinuous on E if 21 is equicontinuous at each point of E. If 21 is equicontinuous at x0, then clearly each AeW is continuous at x0. If E is also a uniform space, 21 is said to be uniformly equicontinuous on £ if for each vicinity N in F there is a vicinity M in E for which (Ax,Ay)eN whenever (x,y)eM and Ae^l.
13. Equicontinuous mappings, and the theorems of Banach and Banach-Steinhaus 169 For linear mappings we have (1) A set $1 of linear mappings A from the topological vector space E into the topological vector space F is uniformly equicontinuous if and only if it is equicontinuous at o, i. e. if and only if for each neighbourhood V of o in F there is a neighbourhood U of o in E for which A(U) cz V for all Ae^i. For given the vicinity Nv of all (yuy2) with yi~y2eV in F, it follows from (x1,x2)eNu, i.e. xl—x2eU, that A(xl—x2) = Axl—Ax2eV. In other words (AxuAx2)eNv for all AeW. The uniform equicontinuity of a collection of linear mappings therefore follows from equicontinuity at o in just the same way as the uniform continuity of a single linear mapping follows from continuity ato. We now prove the theorem of Banach: (2) Let 21 be a collection of continuous linear mappings A from the complete metrizahle space E into the topological vector space F. 21 is equicontinuous if and only if the set ^l(x) of all Ax, Ae$l, is bounded in F, for each xeE. a) The condition is necessary: if V is a neighbourhood of o in F, there is a neighbourhood U of o in E with A(U)a V for all Ae^l. If xepU then AxepV for all Ae^L, so that 2I(x) is bounded. b) The condition is sufficient. Let V be a closed circled neighbourhood of o in F, and let W be another one with W + W a V. We form the set M= f] A{~l)(W). Because of the continuity of A, each A(~l)(W) is closed, and so therefore is M. Further M is absorbent, for 21 (x0) is bounded for each x0eE, and so there is a p>0 with pAx0eW for 00 all Aety', i.e. px0eM. Since M is absorbent and circled, E= (J nM. By Baire's theorem (§4,6.(1)) one of the sets nM, and therefore M itself, contains an open set. Consequently the set M — M contains a neighbourhood U of o, and we have A(U) cz A(M — M) cz W—Wcz V for all AeM. For normed spaces the theorem takes the following form: (2') If tyi is a collection of continuous linear mappings A from the (B)-space E into the normed space F, and if sup\\Ax\\ = M(x)<oo for each xeE, then sup||^||<oo. Ae% The Banach-Steinhaus theorem now follows easily from this theorem of Banach's. (3) Let An be a sequence of continuous linear mappings from the complete metrizahle space E into the topological vector space F. // the sequence of images Anx is bounded for each xeE, and if it is a Cauchy
170 § 15. Topological vector spaces sequence for the points x of a dense subset M of E, then Anx is a Cauchy sequence for each x. If all the Cauchy sequences Anx have a limit Ax in F, then the mapping A defined in this way is linear and continuous. Proof. Suppose that x0eE. By (2), for each circled neighbourhood V of o in F there is a neighbourhood U of o in E with An(U) a V for all n: We choose an xeM with x — x0eU, and write (4) An x0 — Am x0 = (An x0 — An x) + (An x — Am x) + (Am x — Am x0). If n0 is chosen large enough for Anx — Amx to belong to V for all n,m^n0, then each of the three summands on the right-hand side of (4) lies in V, so that Anx0 — Amx0eV+ V + V; Anx0 is therefore a Cauchy sequence. If Ax = \imAnx exists for each xeE, then the mapping A defined in this way is clearly linear. If V is closed, then since An(U) a V for all n, A(U) cz V, so that A is continuous. Banach [3] developed the theory of metrizable vector spaces which we have described here under apparently weaker hypotheses. He considered vector spaces E on which a translation-invariant metric |x,j;| = |x — y, o| = ||x — y\\ is given, which is only required to satisfy a) a„x-*o if a„->0, b) ax„->o if x„->o, c) E is complete with respect to the metric. We now show: (5) A vector space E on which a translation invariant metric is defined, with properties a), b) and c), is a complete metrizable topological vector space. We need only show that £ is a topological vector space when the sets UE9 £>0, are taken as a base of neighbourhoods of o, where UE is the closed ball given by ||x||^£. We shall prove that the conditions of 1.(2) are satisfied. The sets UE clearly form a filter-base with f] UE=o. (LT1) is valid, £>0 1 since UE/2 + UE/2 <= Us, and (LT3) is satisfied since - xe UE always holds for a suitable n, by a), so that xenUE. The main difficulty is to prove (LT2), and so to prove that ax is continuous in both variables; it is only assumed in a) and b) that ax is continuous in each variable separately. We prove this using the method of proof of the theorem of Banach; this itself is not directly applicable to the present situation. For each xeE there is an integer k for which XxeUE for all 1 \k\ ^ —. For otherwise there would be a sequence Aw->0 with |a„x|>e k
14. Bilinear mappings 171 for all n, contradicting a). We denote by M the set of all xeE for which Axe UE for all |A| ^ 1. Since UE is closed, and because of b), M is closed. ][f x is an arbitrary element of F, and if XxeUE for all |/l| ^ —, then X oo /C -eM, so that xekM. Thus E= (J nM, and by Baire's theorem k n= x once again one of the sets kM contains a closed ball Ud(x0). But then M itself contains the ball L^J-yM, for it follows by 12.(5) from ^ - that k ||x0-fcz|l = <k Xq T <s. In other words all the kz belong to Ud(x0)czkM, so that all the z belong to M. M — M contains Us/k(6), so that if |A|^1 and xeUd/k(o) then XxeU2E(o). This establishes (LT2). We observe that the neighbourhoods used in the proof of (5) need not be circled, and therefore ||x|| does not necessarily satisfy (F3); using the results of (5) and 11.(2), however, we can pass over to an equivalent (F)-norm. 14. Bilinear mappings. We defined bilinear mappings in § 9,7. Here we consider bilinear mappings B(x,y) from the topological product Elx E2 of two topological vector spaces into a third topological vector space F. For fixed x, B(x,y) defines a linear mapping Bx from E2 into F, and for fixed y it defines a linear mapping By from El into F. A bilinear mapping B from Elx E2 into F is said to be continuous if it is continuous as a mapping from Ex x E2 into F, and so if it is continuous in both variables together. If this is so, then in particular all the mappings Bx and all the mappings By are continuous linear mappings from E2 and El9 respectively, into F. If it is only assumed that all the mappings Bx and all the mappings By are continuous, then B(x,y) is said to be separately continuous. Equicontinuity and separate equicontinuity are defined in a corresponding way for a collection 95 of bilinear mappings. (1) A bilinear mapping B fa collection 95 of bilinear mappings) is continuous (equicontinuous) if it is continuous (equicontinuous) at (o,o). We prove this for equicontinuity. We write (2) B{x,y)-B{x0,y0) = B{x0,y-y0) + B{x-x0,y0) + B(x-x0,y-y0).
172 § 15. Topological vector spaces Because of the equicontinuity at (o,o), given a neighbourhood W of o and Wl-\-Wl-\-Wicz Wcz F, there is a neighbourhood U x V of (o,o) in El xE2 for which £(x-x0, y — y0)e Wl for all Be95 and all x-x0et/,);-);0eKIf-x0eir, then B(x0,y-y0) = B[—,n(y-y0))eWl n \n J for all B and for all y — y0e—V. Likewise it follows from —y0eV n m 1 that B{x — x0,y0)eWl for all B and all x — x0 e—U. Thus we have (i \ B(x,y) — B(x0,y0)eW for all B, provided that (x — x0,y — y0)e — U) We have the following important theorem (Bourbaki) : (3) Every separately continuous bilinear mapping B(x,y) from the product of two complete metrizable spaces El,E2 into a topological vector space F is continuous. A collection 95 of bilinear mappings B(x,y) is equicontinuous if and only if the mappings B are separately continuous and the set %5{x,y) of all values B(x,y) is bounded in F for each fixed (x,y)eEl x E2. It is enough to show that the conditions of the second assertion are sufficient. We establish a preliminary lemma: (4) A collection 21 of mappings A from a metric space E into a uniform space F is equicontinuous at x0 if whenever x„->x0 *n E tnen Axn converges uniformly to Ax0. Let AT be a vicinity in F. It is enough to show that the set V of all xeE for which (Ax, Ax0)eN for all Ae$l is a neighbourhood of x0. If this were not the case, there would be a sequence x„$ K with x„-»x0, and the sequence Axn would not converge uniformly to Ax0. By (1) and (4) it is therefore sufficient to show that whenever (x„,y„) ->(o,o) in ElxE2 then the sequence B(xn,yn), Be95, converges uniformly to o in F. By hypothesis, for fixed x0 the set 95xo of continuous linear mappings BX0 = B(x0,y\ Be95, from E2 into F maps each y0 into a bounded set. By the theorem of Banach 95xo is therefore equicontinuous; since B(x0,o)=o, for each neighbourhood W ofo in F there is a neighbourhood V of o in E2 with B{x0,y)eW for all Be95 and all yeV. We now show that the set C= \J 95(x0,yJ is bounded in F. «=1,2,... The yn lie in V from n0 onwards, and so [j 95(x0,y„)c: W; for fixed
1. The convex and absolutely convex cover of a set 173 n=l,...,n0 —1, however, 93(x0,y„) is bounded, so that CczmW for some suitable m>0; consequently C is bounded. If we now consider the B(x,y) as mappings By from £x into F, we have shown that the set of mappings Byn, Be95, n=l,2,..., satisfies the hypotheses of the theorem of Banach. This set is therefore once again equicontinuous, and so there is a neighbourhood U of o in £x for which (J <&{x,yn)cz W. For n^nu therefore B{xn,yn)eW for all £e93; xel/ n= 1,2,... this completes the proof. § 16. Convex sets 1. The convex and absolutely convex cover of a set. Convex sets in arbitrary real or complex vector spaces have properties which are essentially more complicated than in the rc-dimensional case. We shall give a brief introduction to the theory of these sets, without striving for completeness. Not everything will be needed later. Reference may be made to the accounts of Bourbaki [6] and Klee [2]. Convex and absolutely convex sets have already been defined in § 15,10. The concept of convex set is the same for both real and complex spaces, while the concept of absolutely convex set is different in the two cases. Because of this, we shall from time to time speak of real or complex absolutely convex sets. We observe that every complex vector space E can also be interpreted as a real vector space; if {xa} is a complex algebraic basis of E then the vectors xa and ixa together form a real algebraic basis of E. If follows directly from the definitions that the intersection of arbitrarily many convex (respectively absolutely convex) subsets is again convex (absolutely convex). The convex cover C (M) of an arbitrary set M is the intersection of all the convex subsets of E containing it. C (C (M))= C (M). If M = (jMa, we shall also write C Ma for C (M). a The (real or complex) absolutely convex cover |~~(M) or \~Ma is defined similarly. n n (1) C (M) consists of all elements of the form Y^0iixh af^0, Za»= *> n xteM. Likewise \~{M) consists of all elements of the form Y,Pixi> PiEK, i
174 § 16. Convex sets Proof. It can be confirmed immediately that the elements of the given form constitute a convex (respectively absolutely convex) set containing M. Conversely, we shall show that these elements belong to every convex (respectively absolutely convex) set containing M. We shall prove this for C (M); the proof for |~~(M) is analogous. n- 1 Suppose that it has been shown for n — 1 that every element £ oct xt n n- 1 i of the above form lies in C (M). Given £ a|xf we put £ ai = a- By the inductive hypothesis y = Y — xf lies in C (M); so therefore does n ! a ay + (l—a)x„ = £ aJ-Xf, since 0<a^l. i= 1 We recall the concept of the circled cover of a set (§ 15,1.). This concept is also different in the real and complex case. (2) The absolutely convex cover of a set M is the convex cover of the circled cover of M. A set is absolutely convex if and only if it is convex and circled. The circled cover of the convex cover of a set M need not be convex. Corollary. If the sets Ma are circled, \~~ Ma= C Ma. a a Proof, a) Since every absolutely convex set is circled, it is sufficient to show that the convex cover of a circled set M is absolutely convex. n n By (1), C (M) consists of all £pf xf w*tn P; = 0, Za = *> anc* so ^ f°ll°ws from (1) that C(M)c=p(M). Conversely given ft + 0 in K with Zlftl^1* if we put ot—pi^-j- and Pi = -—~, we see that every element x = £/?IxIeP(M) can be written in the form Y,Piaixi> w*tn £pj = l, Pi^O, |af|^l a^Xj-eM, so that x belongs to C (M). Hence we also have T(M) c C (M). b) Neither the real circled cover in P2 nor the complex circled cover in l~2 of the convex cover of the three points (0,0), (1,0) and (1,1) is convex (proof!). c) The corollary follows from the fact that [JM^ is circled if the Ma are, and so by (2) we have C Ma= C[\jMa)=r UM* =rM*' (3) // C1?...,C„ are convex (respectively absolutely convex) and a!,...,ocn are arbitrary elements of K, then alCl+-- + anCn is convex (absolutely convex).
1. The convex and absolutely convex cover of a set 175 Since aC is convex or absolutely convex if C is, it is sufficient to establish the assertion for C^ + C^ For convex C1 and C2 this follows from the relation T(x + j;) + (l-T)(x/+y)=[TX+(l-T)x/] + [T^ + (l-T)/] for x, x'eCl9 y, y'eC2, Ofgr^l, and the absolutely convex case is proved analogously. x0 + C is clearly convex if C is. This result, however, does not hold for absolute convexity, for which the point o plays a special role. If C is absolutely convex we say that x0 + C is absolutely convex about x0. (4) The linear image and the linear inverse image of a convex or absolutely convex set are again convex or absolutely convex, respectively. Let A be a linear mapping from the vector space E into the vector space F. A(C) is convex if C is, since A(tx + (1 — x)y) = xAx + (l — x)Ay. Further if M is convex in A(E) and if Ax and Ay belong to M, then A(xx + (l—x)y)eM, so that A(~1](M) is convex. The proof for absolutely convex sets is similar. Now suppose that £ is a topological vector space £[£]. (5) The closure C of a convex or absolutely convex set C in E[X] is again convex or absolutely convex, respectively. We shall prove this for absolutely convex sets. The proof for convex sets is similar. We must show that if x0 and y0 belong to C, then ax0 + /?y0 lies in C too, if |a| + |/?|5n. Given a neighbourhood U of o, let V be a circled neighbourhood of o with K+ V a [/.To x0 and y0 there correspond elements x and y in C with x0 — x e V, y0 — y e V. But then (<xx0 + Py0)-(<xx + Py) = (x(x0-x) + P(y0-y)eV+VczU, so that ax0 + Py0eC. If M is a subset of £[£], the intersection of all the closed convex sets containing M is called the closed convex cover of M. The closed absolutely convex cover is defined in a similar way. (6) The closed convex cover of M is equal to the closure C (M) of the convex cover of M. C (M) is also equal to C (M). The closed absolutely convex cover is equal to \~(M)=[~(M). Once again we shall prove this for absolutely convex sets. By (5), \~ (M) is closed and absolutely convex, and it is clearly contained in all the closed absolutely convex sets which contain M, so that it is the closed absolutely convex cover of M. The last assertion follows
176 § 16. Convex sets from the fact that every closed absolutely convex set containing M must also contain |~"(M). (7) // M is open, then so are C (M) and |~~(M). n n An element of |~~(M) is of the form £afxf, £|af| = a^l, a,=N0. i i There is a circled neighbourhood V of o for which all the sets xt+V n are contained in M. But then every element £a;X; + az, zeV, lies in " / |a-l \ * r(M), since it is of the form £ af I xf + — z I, ze K If M is closed, C (M) and r(M) need not be closed. An example of this is given by the real absolutely convex cover of the closed set M in the plane which consists of the points (—1,0), (1,0) and the y-axis. 2. The algebraic boundary of a convex set. Let C be a convex subset of the vector space E. (1) The set C — C is real absolutely convex. By 1.(3), C — C is certainly convex. If 0<t<1 and z1,z2eC, then t(z1—z2) = z1—(tz2 + {1—t)z1) lies in C — C, since C is convex. Since C-C=-(C-C), it follows from this that C-C is real circled. The assertion now follows from 1.(2). If C is an arbitrary subset of £, we call the real linear subspace of E spanned by all the elements of C—C, i.e. by all the differences of points of C, the real linear space L(C) associated with C. (2) // C is convex, L(C)= [j n(C — C), and if C is absolutely convex, oo n= 1 L(C)= (J nC. n= 1 It follows from (1) that \Jn(C — C) is real absolutely convex. Since further every ax, aeP, lies in \Jn(C — C) if x does, Un(C-C) is a real vector space. The assertion about abcolutely convex sets is proved in a similar way. (3) Let C be an arbitrary subset of E. The smallest real linear manifold M(C) containing C is equal to C + L(C) = x0 + L(C), where x0 is any point of C. If a linear manifold z + H contains the set C, then the linear space H parallel to it contains all the differences xt — x2 of elements of C, and so contains C — C. Hence H 3 L(C). On the other hand, if x0eC then clearly x0 + L(C)^ C. Finally it follows from x = x0 + (x — x0) that x + L(C) = x0 + L(C), for arbitrary xeC. A point x0 of a set C c E is called an internal point of C if C is absorbent about x0 in M(C), i. e. if every straight line through x0 which lies in M(C) contains x0 as an interior point.
2. The algebraic boundary of a convex set 177 When M(C) = E, an internal point of C is called an algebraic interior point of C. The collection of all algebraic interior points is called the algebraic kernel C of C. The algebraic hull Ca of C consists of all those points yeE for which there exists an xeC for which \_x,y)aC ([x,y) denotes the real line segment joining x and y, including x and excluding y). A point of Ca which is not an algebraic interior point of C is called an algebraic boundary point of C. The collection of all the algebraic boundary points of C is called the algebraic boundary of C. A set C is said to be algebraically closed if C = Ca, and to be algebraically open if C = Cl. We remark that an algebraic boundary point of a subset of a topological vector space is always a topological boundary point; Example 2 below shows that the converse is not true, even if the set is convex. In P", every convex set C contains an internal point; indeed, if M(C)=Pm, C contains an m-dimensional simplex, and the centroid of this is an internal point of C. For infinite-dimensional spaces this result does not always hold. Example 1. Let {xp}, /JeB, be an algebraic basis of the infinite-dimensional vector space £, and let C be the convex cover of o and all the xp. M(C) = E. We shall show that C has no algebraic interior points, so that it consists of algebraic n boundary points alone. By 1.(1) every element z of C is of the form z = YJ0LixPi, tfj^O, £ai=l- If P=¥Pi, i=l,...,w, then the straight line z + a(xp — z) through xp and z meets C in the segment [x/j,z], so that z is not an interior point of C. (4) // C is convex, the algebraic hull Ca and algebraic kernel Cl are again convex (the empty set is taken to be convex). Proof, a) If y1 and y2 are two points of Cfl, and if xY and x2 are two points of C with [x^^) cz C and [x2,y2) <= C, then xu x2, y\ and y2 are the four vertices of a tetrahedron, all of whose interior points belong to C. But then the boundary points belong to Ca and so therefore does [yi,y2]- b) Suppose that xx and x2 lie in C, that x is any point between x1 and x2, and that g is any straight line through x. Then on the straight line through xl (respectively x2) parallel to g there is an interval [y^Zj] (respectively [^2)z2]) lying in C. The intersection of g with the quadrilateral yu zu z2, y2 is contained in C; it is an interval in g which contains x as an interior point. Thus x belongs to O. In 4. we shall show that (C^^C1. We now give a counter-example to show that (Ca)a = Ca does not always hold. Example 2. Let E be real and infinite-dimensional, and let {xa}, aeA, be an algebraic basis of E. We denote by C the set of all non-zero x = ^^axa whose coefficients are non-negative and satisfy Yj%* = ~T~\> where n(x) is the number 12 Kothe, Topological Vector Spaces I
178 § 16. Convex sets of a with £a + 0. C is convex, for if z = £(axa = Tx + (l — x)y, y = YJy\ax^ with x and j> in C, then «(z)^max(«(x),«(y))>0; consequently "M n(jO w(z) which shows that zeC. Let us determine Cfl. We assert that Ca is equal to the set of all x = ££axa with £a^0 and w(x)>0. If x = ££axa is any non-zero element with £a^0, then ££a>0, so that there is a positive integer m for which ££a^—. Then element x _|_... _|_x wi z = _*i am jies -n q jf we choose tjje a t0 ^e different from all those indices m oc for which the coefficients £a of x are non-zero, then each tx + (1— t)z with 0^ t < 1 belongs to C. x therefore belongs to C, as it is the end point of the segment [z,x). Further if y is any element of C, the whole of the segment \y,6) is not in C, but only y, , ^— V L so that o does not belong to Ca. L "W&a J Finally osCa\ so that Ca*Ca\ As we have just seen, Ca need not be algebraically closed. We can however repeat the process of forming the algebraic hull transfinitely. We put Ca = Cl\ for an ordinal y + 1 we put Cy+l=(Cy)a, and for a limit ordinal p we put Cp = \J Cy. By (4) all the sets Cy + 1 are convex y<P if C is, and so are the sets Cp, and we must have CM+1 =CM for a sufficiently large ordinal \x. Thus C is the smallest algebraically closed convex set which contains C. (5) Every convex set C has an algebraically closed hull Ca, which is again convex. O. Nikodym [1] has shown that in fact arbitrarily large ordinals \x of cardinality 5^X0 can be needed; cf. Klee [6] as well. The next result gives a further example of the complicated nature of convex sets in infinite dimensional spaces. (6) In every infinite-dimensional vector space E there is a proper convex subset C with Ca = E. Again, let {xa}, aeA, be a real algebraic basis of £, and let the index set be ordered in a way such that there is no final element. Let C be n the set of all £ £fxa.=#o, n = l,2,..., whose last non-zero coefficient is i = l positive. C is clearly a proper convex subset of E. If z = XCjXa. *s an arbitrary point of £ and if a>af for i=l,...,n, then z is the end-point of the segment [xa,z), which lies in C.
3. Half-spaces 179 3. Half-spaces. For the present, let £ be a real vector space. A hyper- plane in E is either a linear subspace H of codimension 1 or a set x0 + H. We know (§15,9.(1)) that corresponding to H there is always a real functional u in the algebraic dual space £* whose null-space is exactly H. The hyperplane x0 + H is then the set of all the xeE for which u(x) = u(x0) = y. The hyperplane x0 + H determines two algebraically open half- spaces, defined by w(x)<yand u(x)>y respectively; similarly it determines two algebraically closed half-spaces, defined by u{x)^y and u(x)^y. It is a simple consequence of the linearity of u that all these half- spaces are convex, that they are respectively algebraically open or closed and that u(x)f^y is the algebraic hull of u(x)<y. The hyperplane x0 + H is the algebraic boundary of each of the four half-spaces. A further consequence of § 15,9.(1) is that, in a topological vector space E[X~\, the half-space u(x)^y is closed if u is continuous. If this is the case, the half-space u(x)<y is open. If u is not continuous, all four half-spaces are dense in £[2]. If £ is a complex vector space, £* consists of complex linear functional on E, and the question arises of how to characterize the half- spaces of a real hyperplane by means of a complex linear functional. (1) If a complex vector space E is considered as a real vector space, and if u is a real linear functional on it, then there is a uniquely determined complex linear functional v on the complex vector space E for which ux = y{(vx). If u is continuous on a complex topological vector space £[3T], then v is also continuous; conversely, it follows from the continuity of the complex linear functional v that the real linear functional ux = 9i(vx) is continuous. Proof. If ux is to be equal to the real part of ux, then the imaginary part 3(i;x) must satisfy 3(i;x) = 9?( — ivx)= — <R(v{ix))= — u(ix). The only possible extension to v is therefore (2) vx = ux — iu(ix). This is certainly a complex-valued linear function on E, when E is considered as a real space, v is however also complex linear, for we have v(i x) = u(i x) — iu( — x) = i[ux — iu(i xf\ = ivx. If |wx|^e for all x belonging to a complex circled neighbourhood U of o in E[X~\, then by (2) \vx\^2e on U, so that v is continuous if u is. Conversely it follows from |ux|^£ that |9t(i;x)|^£. 12-*
180 § 16. Convex sets The analytic characterization of real hyperplanes x0 + H and half- spaces in a complex space E follows from (1): corresponding to x0 + H there is a complex linear functional veE* for which the points of x0 + H are characterised by the equation 9i(vx) = 9i(vx0) = y. The algebraically open half-spaces are given by <H(vx):^y. These half- spaces are topologically open if and only if v is continuous. Every real hyperplane H through o contains just one complex hyper- plane, namely HniH. If H is given by ux = 0, HniH is given by vx = 0, where v is defined by (2). 4. Convex bodies and the Minkowski functionals associated with them. A particularly important class of convex sets is formed by those which have at least one algebraic interior point. These are called convex algebraic bodies, or convex a-bodies. The paradoxical possibilities of 2. cannot arise for such convex sets, as we shall see. Every absorbent absolutely convex set is a convex a-body, since o is an algebraic interior point. If the underlying space is a topological vector space £[£], we call a set Ca convex I-body if C has an interior point x0, in the sense of the topology X, and so if C is a convex ^-neighbourhood of x0. Every convex £-body is a convex a-body, but not conversely. If C c E is a convex a-body with o as an algebraic interior point, we define (cf. § 15,10.) the Minkowski functional q(x) by the equation (1) q(x)=infp (p^O) xepC (2) The Minkowski functional of a convex a-body C with o as an algebraic interior point is a non-negative, positive homogeneous, subadditive function on E; i. e. for all x, yeE it satisfies the conditions (P) q{ox) = oq{x) for a^O, Proof. Because C is absorbent, q(x) is defined for all xeE, and (a) and (/?) are clearly satisfied, (y) is proved as in § 15,10: Let q(x) = p0, x y q(y) = G Then if p>p0 and g>g0, the elements - and - he in C, p g t . ^ . 1 p x g x x+y and since C is convex, so does 1 = . Hence P + G p p + G G p + G x + ye(p + o-)C, so that q(x + y)^p0 + G0.
4. Convex bodies and the Minkowski functionals associated with them 181 It follows from (/?) and (y) that q(x) is a convex function on £, i. e. (3) q(rx + (l -t)jO ^ t^) + (1 -t)^), O^t^ 1, for all x,yeE. (4) // g(x) is the Minkowski functional of the convex a-body C with o as an algebraic interior point, then O consists of all x with q(x)<\, and Ca consists of all x with q(x)^l, so that the algebraic boundary of C is given by q(x)=\. Further, O is an algebraically open convex a-body, and Ca is algebraically closed. Proof. By (1), q(x)^\ if xeO If q(x)<\, then, conversely, xeC. We now show that such an x also lies in O. Let q(x) = x<\, and let 0<(7<1— t. The convex a-body x + oC has x as an algebraic interior point. It follows from q{x + <jz) ^ q(x) + aq(z) ^ t + (7<1, for zeC, that x + aCaC, so that every point x with q(x)<l lies in O. Indeed, every such point lies in (Of, for it follows from q(x + oz)<\ that x + aCcC1, so that x is an algebraic interior point of O. Hence O is also a convex a-body. If q(x) = \, then because of (/?), the interval [o,x) lies in C, so that all those points with q(x)f^l lie in Ca. Conversely, suppose that xeCa. There exists an interval [z,x)aC. Then since z + t(x — z)eC for all 0^t<1, it follows from x = z + x(x — z) + (1 — x)(x — z) that ^[(x)^^(z + t(x-z)) + (1-t)^(x-z)^ l+(l-T)g(x-z). Letting t->1, it follows that q(x)^\. Ca is therefore characterized by^W^l. If q(x)=\, then x does not belong to O, for otherwise a segment [o,(jx], g> 1, would lie in C, and this is not possible, by (2) (/?). Consequently, (Of and C only contain points for which q(x)<\, and so consist of the collection of all such points; O is therefore algebraically open. Finally, if the argument which was given in the preceding paragraph for C is applied to Cfl, we obtain (0)a = 0, so that O is algebraically closed. We observe that it also follows from (4) that (O)a = 0. Since [o,y~\ czC whenever yeO, every algebraic boundary point of C can be reached from every algebraic interior point by a straight line in C. We have the following converse:
182 § 16. Convex sets (5) // q(x) is a function on E which satisfies conditions (a), (/?) and (y), then the set q(x)< 1 (respectively q(x)^ 1) that it defines is an algebraically open (respectively algebraically closed) a-body with o as an interior point, whose Minkowski functional is q(x). Proof. The set C of all x with q(x)^l is clearly absorbent, o is an algebraic interior point of C. It follows directly from (3) that C is convex. The Minkowski functional of C is q(x) again, and the remaining assertions follows from (4). We remark that if q{x) = 0 for some non-zero x, this means that the whole of the half-line from o through x is contained in C. In particular E itself is described by the functional q(x) = 0. Starting from a convex body C, and forming its Minkowski functional, we can certainly recover O and Ca, but cannot recover C itself, as C can be any arbitrary convex set satisfying C' cz C cz Ca. If we make the additional hypothesis on C that it is an absolutely convex subset of the real (respectively complex) vector space E, then in place of (/?) we obtain the relation q((xx) = \a\q(x), for arbitrary real (respectively complex) a. In this case, therefore, q(x) is a real (respectively complex) semi-norm (cf. § 14,1.). Consequently, we have (6) The Minkowski functional of an absorbent absolutely convex set is a semi-norm, and conversely. The question of when a convex a-body in a topological vector space E[Z~] is a £-body is answered by (7) A convex a-body C in E[X~\ with o as an algebraic interior point is a X-body if and only if its Minkowski functional is continuous. If this is so, O is the interior of C and Ca is the closure of C; the algebraic and topological boundaries of C are therefore the same. Proof, a) Suppose that o is an interior point of C. Then q(x)^e for all x in the neighbourhood eC of o, i. e. q is continuous at o. It follows from the continuity of q(x) at o, however, that q(x) is continuous at an arbitrary point yeE, and indeed that q(x) is uniformly continuous: If zeeV, then q(y + z)f^q(y) + s; similarly, the inequality q(y)-zi%q(y + z) follows from q(y) = q((y + z) — z)^q{y + z) + q(-z), since —zecV. Hence \q(y + z) — q(y)\z%s for zezV. b) Conversely, if q(x) is continuous, the set of all x with q(x)<\, i.e. C, is open and contains o, and the set q(x)^\, i.e. Ca, is closed, so that O is the interior of C and Ca is the closure of C. The geometric properties of convex bodies with o as an interior point naturally also hold for convex bodies with a general point x0 as an interior point. If x0 is an algebraic interior point of C and if q(x) in the Minkowski functional of C — x0, then Ca is given by q{x — x0)^l.
5. Convex cones 183 We now make some comments indicating how these ideas can be applied to more general convex sets. If C is a convex set for which O is empty, then (C)1 is also empty, so that we obtain the following general result from (4): (8) // C is an arbitrary convex set, then (Cl)l = Cl. If C is an arbitrary convex set, with corresponding linear manifold M(C), and if C has at least one internal point, then C is a convex a-body in M(C), and we obtain from (4) that the set of internal points is algebraically open in M(C), and that (Ca)a = Ca. In particular, this result applies to all the convex subsets of an n-dimensional space, since these have internal points, by (2). If a convex set has at least one internal point, and if Ca = M(C), then it follows from (4) that we must have C = M(C), and the pathological behaviour of example 2.(6) is therefore no longer possible. 5. Convex cones. A subset K(x0) of a vector space E is called a cone with vertex x0 if K(x0) contains every point x0 + p(x — x0), p>0, whenever it contains x. A cone with vertex o therefore contains px,p>0, whenever it contains x. A cone K(x0) can always be obtained as a translate x0 + K(o) of a cone K(6) with vertex o. A cone K(o) is convex if K(o) always contains x + y when it contains x and y. Conversely if a convex cone K(o) contains x and y it always contains Ax + fiy, for arbitrary positive / and \i. If K(o) is a cone, so is —K(o). More generally, x0 — K(o) is called the cone diametrically opposite to K(x0) = x0 + K(o). We denote it by K*(x0). The real vector space L(K(x0)) corresponding to a convex cone K{x0) = x0 + K(o) is equal to K(o)-K(o), by 2.(2). A cone is said to be proper if it contains no real line through its vertex. It is said to be truncated if it does not contain its vertex, and to be pointed, if it does. A truncated cone is always proper. By removing the vertex of a proper pointed cone, a truncated cone is obtained, which is convex if the original cone was. Conversely, if the vertex is added to a truncated convex cone, then a proper pointed convex cone results. Acone X(o) is proper if and only if either K(o)nK*(o)=o or X(o)niC*(o) is empty. Every linear manifold is a convex cone. An algebraically open half-space is a truncated convex cone, and any point of the boundary hyperplane can be chosen as vertex. The convex set C of 2. (6) is a truncated convex cone with vertex o which has no algebraic interior point. If K(6) is a cone, so is {K(oj)a. The example which we have just mentioned shows that if K(o) is a proper cone, {K(o))a need not be proper.
184 § 16. Convex sets The linear image and the linear inverse image of a cone with vertex o are again cones with vertex o. The intersection of cones all with the same vertex is again a cone with the same vertex, and the same holds for the union. (1) Given a collection Ka(o) of convex cones, the smallest convex cone containing all the Ka(o) is equal to £ Ka(o). For X^a(°) is a convex cone, since the sum of two elements in a Y.K^o) belongs to £Ka(o), and it is clearly the smallest convex cone a a containing all the cones Ka(6). The cone with vertex x0 generated by a set M is the smallest cone with vertex x0 which contains all the elements of M. (2) // M is convex, the cone with vertex o generated by M is convex, and is equal to [j pM. If o $ M the cone generated by M is truncated, p>0 and is therefore proper. The set [j pM is a cone, and it contains Ax + jny if A and \i are p>0 positive and x and y belong to it. The vertexo is not contained in [j pM if o does not belong to M. p>o If we construct the set C of example 2 of No. 2 in some hyperplane of a vector space and use it to generate a cone with vertex outside the hyperplane, then we obtain a convex cone for which (Ka)a=^Ka. Convex cones therefore exhibit the same pathological properties as arbitrary convex sets. The example of the convex set £vj ^ 1, £, r\ > 0, and the cone with vertex o which it generates shows that the cone generated by a closed set need not be closed. In topological vector spaces, we have (3) // K is a convex cone in E[X~\, then so also is its closure K. Kl is a convex cone as well, and Kl is the interior of K, if Kl contains an interior point. Proof. We can take the vertex to be o. By 1.(5), K is convex. If z is a closure point of the points xeK, then pz, p>0, is closure point of the points px, which also lie in K, so that K is a cone. If Kl is empty, then the second assertion is true. If Kl is not empty, then K is respectively a convex algebraic body or a convex £-body, and the second assertion follows from 4.(4) and 4.(7), since Kl is again a cone. 6. Hypercones. A maximal convex truncated cone in E with vertex x0 is called a hypercone at x0. We have the following important existence theorem: (1) // M is convex, and if x0 does not belong to M, then there is a hypercone at x0 which contains M.
6. Hypercones 185 Proof. By 5.(2) there is a truncated convex cone with vertex x0 which contains M. Since the union of a totally ordered collection of such cones (ordered by inclusion) is again truncated and convex, it follows from Zorn's lemma that there is a maximal such cone. The diametrically opposite cone K*(x0) is a hypercone if K(x0) is, and we have (2) // K(x0) is a hypercone, then K(x0)uK*(x0) = £^{x0), and K(x0)nK*(x0) is empty. The complement of a hypercone K(x0) is therefore the convex cone K*(x0)u {x0}. The assertion that K(x0)nK*(x0) is empty is obvious, since a hypercone is convex and truncated, and is therefore proper. The first assertion follows from maximality: we may clearly suppose that x0=o. If Kkj(-K) were not equal to E~{o}, there would be a real line ax not lying inXu(-K). But then the collection of points px with p>0, together with the points px-\-y with yeK and p^O, would form a truncated convex cone (since px + y=to), and this would contain K as a proper subset, which is not possible. (3) Conversely, if K(x0) is a convex cone with X(x0)uK*(x0) = E~{x0} and with K(x0)nK*(x0) empty, then K(x0) is a hypercone. The significance of hypercones in the study of convex sets results from (4) Every proper convex subset CofE is the intersection of the hypercones containing it. For if x0$C, there is a hypercone at x0 containing C and not x0, by (1). It follows immediately from (3) that (5) The intersection of a hypercone K(o) with a linear subspace H of E is a hypercone in H. Example 2.(6) is an example of a hypercone whose algebraic hull is equal to E. If we apply Theorem 2 of § 17, 1., then in particular we obtain from (2) that the algebraically closed hull of a hypercone is either E or an algebraically closed half-space whose bounding hyperplane is the algebraic boundary of the hypercone. In the latter case the hypercone is a convex a-body. In the former case the hypercone has no algebraic interior point (cf. the final remark of 4.). A hypercone K{o) in P" is a convex a-body, so that it has a hyperplane as algebraic boundary. The points of K{o) lying in this space again form a hypercone, and so on. It is an easy consequence of this that a hypercone in P" always has the following form: there is a basis xl9...,x„ of P" such that K(o) consists of all the n points Yj A\x&° f°r which the last non-zero co-ordinate is positive. i= 1
186 § 17. The separation of convex sets. The Hahn-Banach theorem § 17. The separation of convex sets. The Hahn-Banach theorem 1. The separation theorem. Once again, let E be a vector space over K, where K is the real or complex field. (1) // A1 and A2 are two disjoint convex proper subsets of E, there exist two complementary convex subsets C1 and C2 of E with Cx Dib C2=> A2. We shall give two proofs of this important theorem (cf. Hammer [1] and Bourbaki [6], Vol. 1, p. 53). a) Ax— A2 is a convex subset of E, by § 16,1.(3), and it does not contain o. By § 16, 6.(1), there is a hypercone K with vertex at o for which K^Al-A2. We set Cx= f] (x2 + K). Since Ax-x2^K, so that x2eA2 Ax c= x2 + K, it follows that Al<=Cl. By § 16,6.(2), the complement of a set x2 + K is equal to x2 + (K*u{o}), so that the complement of Ct is equal to the union C2 = A2 + (K*u {o}) of these complements. C2 is convex, by § 16,1.(3), and C2 => A2. b) The following proof is independent of the theory of hypercones. We consider the collection of pairs (B1,B2) of disjoint convex sets with Bx => Al and B2 => A2. The collection of these pairs is partially ordered if we put {BUB2)^{B'UB'2) when B1czB\ and B2<=B'2. Suppose that BlvB2 + E. We shall show that another pair (B\,B2) exists with (B1,B2)<(B'1,B'2). Suppose that x0^B1kj B2. We assert that either the convex cover of x0 and Bx has an empty intersection with B2, or the convex cover of x0 and B2 has an empty intersection with Bt. If this were not so, then for some y1eBl there would be a point z2eB2 in the segment [x0>J>i]> and for some y2eB2 there would be a point z1eBl in the segment [x0,j/2]. But then the point of intersection of the segments [j^,zx] and [^2^2] would lie in both B1 and B2, giving a contradiction. The result now follows by applying Zorn's lemma to the collection of pairs (Bl,B2). (2) // Cx and C2 are proper complementary convex subsets of E, then C\ n C2 is either equal to E or is equal to a real hyper plane. In the latter case, i. e. if C\ and C2 are not both empty, C\ and C2 each coincide with one of the two algebraically open half-spaces defined by this hyperplane. Proof. Weset Ca1nCa2 = H. E-H = C\uC2, for since C1kjC2 = E, each boundary point of Cx (respectively C2) is also a boundary point of C2 (respectively CJ, and so lies in H. For the same reason, H is not empty. CJnC2 is convex, by §16,2.(4). Further, if H contains two points zx and z2, it contains the whole of the straight line through them. For suppose that this were not the case, and that z was a point of the line, lying outside the segment \_zx,z2\ which did not belong to H. Let us
1. The separation theorem 187 suppose that z2 lies between z and zx. z would lie either in C\ or in C2. Suppose that zeC\. Then it follows from the remark preceding § 16, 4. (5) that every point of the segment (zl,z~\ would lie in C\, so that z2 would, and this is impossible. Thus we have shown that H is a linear manifold. Now suppose that H#=E. We can suppose that oeH, so that H is a linear subspace. Suppose that x0$H, and that x0eC\, say. Then — x0 does not belong to H, either, and so — x0 e C\ u C2. But — x0 cannot belong to C\, for otherwise, o, being a point of the segment [x0, — x0], would lie in C\, since C\ is convex (§ 16,4.(4)). Thus — x0 belongs to C2. We assert that H(B\_x0~] = E, where [x0] is the space of all real scalar multiples of x0. If xeCl9 [x, — x0] contains a point of H, so that Clcz H©[x0]; if yeC2, [y,x0] contains a point of H, so that C2 c= H © [x0], as well. Consequently H © [x0] = E. Since E~H = C\uC2, one of the algebraically open half-spaces must concide with C\, and the other with C'2. We say that two sets M and N are separated by a real hyperplane H if they are contained in different algebraically closed half-spaces defined by H. We also say that they lie on opposite sides of H. M and N may have points of H in common. M and N are said to be strictly separated by H if they are contained in different algebraically open half-spaces defined by H. The algebraic form of the separation theorem follows from (l)and(2): (3) // Ax is a convex a-body and A2 is a convex set which contains no algebraic interior point of Al9 then there is a real hyperplane H which separates Ax and A2, and which contains no algebraic interior point of Ax. If A1 and A2 are disjoint algebraically open convex a-bodies, then there exists a strictly separating real hyperplane. Proof. By hypothesis A\ is non-empty and A\nA2 is empty. By (1), there exist complementary Cl^>A[, C2^>A2. Since C\ is nonempty, there exists a hyperplane H, by (2), for which C\, and therefore A\, lies in one of the algebraically open half-spaces and A2 lies in the complementary algebraically closed half-space. This establishes the first part of the theorem. If Al = A\ and A2 = A2, then, since A2<=Cl2, A\ is contained in the other open half-space; in this case, therefore, the separation is strict. In a topological vector space E[X~\ the separation theorem takes the following geometric form (cf. M. Eidelheit [1], S. Kakutani [1]): (4) // Ax is a convex %-body in £[£], and if A2 is a convex set containing no interior point of Al9 then there is a closed real hyperplane H separating Al and A2, which contains no interior point of Ax.
188 § 17. The separation of convex sets. The Hahn-Banach theorem // Ai and A2 are disjoint convex open X-bodies, then there exists a strictly separating closed real hyperplane. Proof. H is either dense in E or closed, by § 15, 9.(1). Since the set A\ is open (§ 16,4.(7)) and disjoint from H, H must be closed. It is not always possible to separate two disjoint algebraically closed convex a-bodies At and A2 strictly: in the plane, take Ax to be the half-plane £^0, and take A2 to be the set of all (£,*/) with £rj^l, £,ri>0. 2. The Hahn-Banach theorem. By considering a special case of the algebraic form of the separation theorem we obtain (1) If C is a convex <x-body in a vector space E and if M is a linear manifold which contains no algebraic interior point of C, then there is a hyperplane H containing M which again contains no algebraic interior point of C. For real E, (1) follows as a special case of 1.(3), when we replace the hyperplane separating C and M by a parallel hyperplane through a point of M; this must contain the whole of M. If E is complex, there is a real hyperplane H with the required properties, again by 1.(3). It is sufficient to consider the case where M goes througho. But then HniH is a complex hyperplane (cf. § 16, 3.) which again contains M, since MniM = M, and a fortiori it contains no interior point of C. From the geometric form of the separation theorem we obtain the following theorem, originally established by Mazur [2], but called the geometric form of the Hahn-Banach theorem by Bourbaki : (2) // C is a convex %-body in a topological vector space £[£], and if M is a linear manifold which contains no interior point of C, then there is a closed hyperplane H which contains M, and which again contains no interior point of C. If we apply the analytic descriptions of convex bodies and half- spaces which were established in the preceding section, (1) and (2) can be expressed in analytic form. We must now distinguish between the real and the complex case. (3) Suppose that a non-negative positive-homogeneous sub-additive function q(x) is given on a real vector space E. If a linear functional /(z), defined on a linear subspace F, satisfies (4) Kz)£q(z) for zeF, then l(z) can be extended to a linear functional u, defined on the whole of E, which satisfies (5) ux^q(x) for xeE. If E is a topological vector space and q(x) is continuous, then ux is also continuous.
3. The analytic proof of the Hahn-Banach theorem 189 Proof. By § 16,4.(5), the inequality q(x)< 1 defines an algebraically open convex a-body Cbo. We assume that l(z) does not vanish on the whole of F (otherwise the linear functional which vanishes identically is a solution for (5)). There exists a point z0 e F with l(z) = 1, and F = [z0] © Ft, where / vanishes identically on Flt / is therefore identically one on the linear manifold z0-\-F1. By (1), there exists a hyperplane z0 + /f, with Ft a H, which contains no point of C. z0 does not lie in H, so that each point x of E has the form x = az0 + y, yeH. We define u by ux = u(az0 + y) = (x. u is an extension of /, and again u takes the value 1 on the whole of z0 + H. The body C lies in the half-space u x < 1 to which o belongs. It is now easy to see that inequality (5) holds: since q is positive- homogeneous, it is sufficient to prove (5) for one point of each ray through o. If a ray cuts z0 + H at a point x, then the most x can be is a boundary point of C, i.e. q(x)^l=ux. A ray which does not cut z0 + H only contains points for which wx^O, so that (5) is satisfied, since g(x)^0. Finally, if q(x) is continuous, C is a convex £-body, and so the linear functional u defining H is continuous, by § 15, 9.(1). For complex vector spaces we obtain (6) Suppose that a non-negative positive-homogeneous subadditive function q(x) is given on a complex vector space E. If a complex linear functional /(z), defined on a complex linear subspace F, satisfies (7) 9ll(z)£q(z), for zeF, then l(z) can be extended to a complex linear functional v defined on the whole of E, which satisfies (8) Mvx^q(x), for xeE. If E is a topological vector space and q(x) is continuous then v is also continuous. This follows from (3), since 9t/(z) is a real linear functional on the space F, considered as a real vector space; it can be extended to a linear functional which, by § 16, 3.(1), can be written uniquely as the real part of a complex linear functional vx. The continuity of v follows likewise from the continuity of 9iv. 3. The analytic proof of the Hahn-Banach theorem. We have given two essentially geometric proofs of the Hahn-Banach theorem, which follow from properties of convex sets. The classical proof (cf. Helly [1], [2], Hahn [2], Banach [2]) is of an analytic nature, and does not use convexity. It provides a somewhat more general result than 2.(3), as the hypothesis that g(x)^0 can be omitted.
190 § 17. The separation of convex sets. The Hahn-Banach theorem (1) (Hahn-Banach theorem). Suppose that q(x) is a positive- homogenenous subadditive function on a real vector space E. If a linear functional /(z), defined on a linear subspace F, satisfies (2) Kz)^q(z), for zeF, then l(z) can be extended to a linear functional u, defined on the whole of E, which satisfies (3) ux^q(x), for xeE. If E is a topological vector space and q(x) is continuous at o, then u is also continuous. Proof. Suppose that l(x) is defined for Fx => F, and that (2) is satisfied on Fv We shall show that if x0iFl the linear functional l{x) can be extended to [x0] © Fx in such a way that (2) is still satisfied. If z and z are arbitrary elements of Ft, then because of (2) and the properties of q(x) we have l(z')- l{z) = l{z'-z)Sq[(z' + x0) + (-z-x0)] ^q(z' + x0) + q(-z- x0), so that -^-z-x0)-/(z)^9(z' + Xo)-/(z'). Since this holds for each z, z'eFl9 we have sup[-g(-z-x0)-/(z)]^ inf [4(z' + x0)-/(z')]. zeFi z'eFi Let y be a number lying between these two values, i. e. let (4) -q(-z-x0)-l(z)^y^q(z + x0)-l(z) for all zeF,. We now set l(ax0 +z) = ay+ l(z) for all zeF1. In this way / is extended to [x0] © Fx. We shall show that (2) is satisfied. First suppose that a > 0. It follows from the right-hand side of (4) that y^qi—l-x0l—/( — ), and so that ay^aqi hx0|—a/( — = g(z + ax0) — /(z); i.e. /(z + ax0) = ay + Z(z)^g(z + ax0). On the other hand suppose that a= — p, p>0. It follows from the left-hand side of (4) that -q[ x0 I + /( - J ^ y. Multiplying by p, we get — q(z + otx0) + l(z)^py, and so /(z + ax0) = ay + /(z)!g#(z + ax0). The existence of u follows either by repeating the process, using transfinite induction, or by using Zorn's lemma. If q(x) is continuous at o, then q(x)^s for all x in some circled neighbourhood U of o. It
3. The analytic proof of the Hahn-Banach theorem 191 follows from ux^q(x)^e and — ux = u( — x)^q( — x)^e that |ux| ^e for xelJ, and this implies that u is continuous. The complex case results from this, as in the preceding number. It is identical with 2.(6) except that the hypothesis that q(x)^0 is omitted. Conversely, the geometric form 2. (2) of the Hahn-Banach theorem can also be derived from the analytic form (1) or 2.(3): let q(x) be the Minkowski functional of a convex a-body C with o as an algebraic interior point. C is given by q{x) < 1. M has the form x0 + F, where F is a real vector space. q(x0 + y)^ 1 for all x0 + y ex0 + F. We set l(<xx0 + y) = <x on [xo]0F. If a>0, /(ax0 + y) = a/( x0 + —) = <x-\^<xqlx0 H \ = q(<xx0 + y). If a<0, /(axo + y)<0, so that ){txx0 + y) <g(ax0 + y), since q(x)^0 on E. Hence the hypotheses of (1) are satisfied, and so there is an extension u of / with uxf^q(x), u(x0)=\ and uy = 0 for yeF. Consequently C is contained in the half-space wx<l, and x0 + F is contained in the hyperplane ux = l. This again proves 2.(1) in the real case; the complex case and 2. (2) follow as in 2. For the most part, the Hahn-Banach theorem is not applied in its general form (1), but in a more special form, which is the same for real and complex vector spaces: (5) Let p(x) be a semi-norm on a vector space E. // l(z) is a linear functional on a linear subspace F which satisfies (6) |/(z)|^p(z), for zeF. then l(z) can be extended to a linear functional u, defined on the whole of E, which satisfies (7) \ux\^p(x) for xeE. If E is a topological vector space and p(x) is continuous, then u is also continuous. Proof. In the real case it follows from l(z)^p(z) on F that there is an extension u with ux^p{x) on the whole of E. But then we also have — ux = u( — x)^p( — x) = p(x), i.e. (7) holds on the whole of E. In the complex case, let us put l1(z) = <Rl(z). Then \ll(z)\^p(z) on F. From what we have just shown, there is a real linear extension ux of li with lUiXl^pix) on the whole of E. By § 16, 3., ux = u1x — iul(ix) is a complex linear extension of l(z) to the whole of E. For arbitrary xeE, let ux = re1^. Then \ux\ = e~i&ux = u(e~i&x) = ul(e~i*x)Sp{e~i^x) = p(x), so that (7) is satisfied. The next result is a special case of (5) which is frequently used: (8) // p(x) is a continuous semi-norm on £[£] and if x0 is an arbitrary fixed point of E, then there is a continuous linear functional u on E with \ux\^p(x) and ux0 = p(x0).
192 § 17. The separation of convex sets. The Hahn-Banach theorem For l(ax0) = ap(x0) defines a linear functional on the one-dimensional subspace [x0] of E, and (5) can be applied to this. Corresponding to (1), we have (9) // q(x) is a positive-homogeneous subadditive function on a real (respectively complex) vector space E and if x0eE, then there is a real (respectively complex J linear functional u on E with ux^q(x) on E and ux0 = q(x0) (respectively with ^fi(ux)^q(x) and <R(ux0) = q(x0)). If E is a topological vector space, u is continuous if q(x) is continuous ato. Proof. In the real case, a linear functional on [x0] is defined by l(ax0) = aq(x0). If a^O, l(ocx0) = q(x0) it follows from q(o) = pq{o) for p>0 that q(o) = 0. Since 0 = q{o)<,q(x0) + q(-x0), -q(x0)^q(-x0), and so if a <0 we have l{ctx0) = ccq(x0)<, -aq{-x0) = q((xx0). The linear functional / defined on [x0] therefore satisfies the hypotheses of (1). It is easy to derive the complex case from the real case. Let us remark that (5) need not hold if a complex-valued linear functional is only defined on a real linear subspace; cf Bohnenblust and Sobczyk [1]. 4. Two consequences of the Hahn-Banach theorem. In §15,9. we saw that there are topological vector spaces on which the only continuous linear functional is the one which vanishes identically. We have the following criterion: (1) There exist non-trivial continuous linear functional on a topological vector space E[%~] if and only if there is a convex neighbourhood ofo in E which is different from the whole space E. Proof. If u is a non-trivial continuous linear functional on E[%~\, the set of all x with |ux| ^ 1 is an absolutely convex neighbourhood ofo in E which is different from E. Conversely, if U is a convex neighbourhood of o the corresponding Minkowski functional q(x) is continuous, by § 16,4.(7). If x0 is a point for which g(x0)4=0, then it follows from 3.(9) that there exists a continuous linear functional on #[£] which does not vanish identically. It follows from (1) and § 15, 9.(9) that the spaces LP, 0<p< 1, have no convex neighbourhood ofo other than LP itself. It sometimes happens that we do not have a linear functional defined on a linear subspace of E, but only a function l(z) defined on a subset of E. We shall restrict our attention to the case corresponding to 3.(5).
5. Supporting hyperplanes 193 (2) Let p(x) be a continuous semi-norm on £[£]. A function l(z) defined on a subset M can be extended to a continuous linear functional u on E\1L\ which satisfies \u(x)\^p(x), provided that the inequality (3) E^o?*) < p(Zafczfc) holds for all n, all zkeM and all real (respectively complex) <xk. Proof. Let F be the linear span of M. If we define a linear functional on F by setting /I £akzfc )==]Tafc/(zfc), this definition is unambiguous, since it follows from Xafczk = 0 that £ak/(zk) = 0, by (3). But the linear functional defined on F in this way satisfies the inequality |/(z)|^p(z); the assertion follows from 3.(5). 5. Supporting hyperplanes. We continue with the study of convex sets. If N is a subset of a vector space £, a real hyperplane H is called a supporting hyperplane of N if H contains at least one point of N and N lies in one of the two algebraically closed half-spaces defined by H. A point of N through which a supporting hyperplane passes is called a point of support of N. (1) IfCis an algebraically closed convex oc-body in E, every boundary point is a point of support of C. If #[£] is a topological vector space, every point of the boundary of a closed convex X-body C is a point of support of a closed hyperplane, and every supporting hyperplane of C is closed. Proof. The first assertion is an immediate consequence of 2.(1), since every boundary point of C can be taken as M. If C is a £-body, every supporting hyperplane must be closed, by §15,9.(1), since it cannot be dense in E. It follows from (1) that (2) An algebraically closed convex a-body C is the intersection of the algebraically closed half-spaces which contain it and which are defined by its supporting hyperplanes. If C is a closed convex %-body in £[£], C is the intersection of the closed half-spaces which contain it and which are defined by its closed supporting hyperplanes. For if y$C and x0 is an algebraic interior point of C, there is a boundary point z0 on the segment between x0 and y. A supporting hyperplane through z0 cuts the straight line through x0 and y in z0, and y belongs to the algebraically open half-space which does not contain C. 13 Kothe, Topological Vector Spaces I
194 § 17. The separation of convex sets. The Hahn-Banach theorem Since a convex set in P" is either a convex body in P" or a convex body in some manifold of P", every closed convex set is the intersection of the closed half-spaces containing it. We now investigate the extent to which (2) can be carried over to arbitrary algebraically closed convex sets. (3) // E has a countable real algebraic basis, every algebraically closed convex subset C of E is the intersection of the closed half-spaces containing it. First we show: // o$C, there exists a convex a-body C, with o as an algebraically interior point, which is disjoint from C. If xl,x2,... is a basis of E, let En denote the linear subspace spanned by x!,...,xn, and let Cn be the convex algebraically closed intersection of C and En. There exists a compact convex subset C\ oiEx which contains o as interior point and which is disjoint from Cx. Applying § 15, 6.(9) to the two-dimensional space E2 with its usual topology, there exists a convex compact neighbourhood U of o in E2 for which (C\ + U)nC2 is empty. But C2 = C\ + U is a convex compact body in E2, by § 15,6.(8) and § 16,1. (3). Repeating this procedure, we obtain a sequence C\ c C2 <= • • • 00 of convex compact subsets of En, whose union C = [j Cn is a convex n= 1 a-body in E with o as an interior point. C r\C is empty. Thus if x$C there is a convex a-body C, with x as an interior point, which is disjoint from C. The assertion now follows by applying the algebraic form of the separation theorem. (3) is false for every vector space with an uncountable basis, as was shown by Klee [2], III: (4) If E is any vector space with an uncountable real basis, there exists an algebraically closed convex set C which does not contain o but which has a non-empty intersection with every convex a-body which contains o as an interior point. In particular, every algebraically closed half-space which contains C also contains o, so that C is not the intersection of the algebraically closed half-spaces which contain it. Proof. Let xa, aeA, be a real algebraic basis of E. We denote 1 " by M the set of all elements —^ £ *ai> n=\,2,..., where the xai are n i = i any n basic elements. Let C be the convex cover of M. The elements of C have non-negative coordinates, and are all different from o. In every linear subspace F spanned by finitely many xa, the intersection CnF is the convex cover of finitely many points and is therefore a closed polyhedron. It follows from this that C is algebraically closed.
5. Supporting hyperplanes 195 Now let C be a convex a-body with o as an interior point. For each xa there is a positive integer k for which -xaeC. Since A is uncount- 1 able, there exists a k for which -xa.eC for a countably infinite set af. Ml 1 fc If we choose k of these xa., then ^ - • - xa.: = -^ X! xai *s ^n C', since i = 1 »v /v rv i C is convex; it is also in C, so that Cr\C is non-empty. If C is a closed convex subset of a topological vector space, C need not be the intersection of the closed half-spaces containing it: in LP, 0<p<l9 there is no closed hyperplane, by § 15,9.(10). The statement of (2) has no useful application to LP, since there is no convex I-body in LP other than LP itself, by 4. There are certainly closed convex sets in LP though, e. g. sets of the form [o,x], xeLp, and none of these can be expressed as the intersection of closed half-spaces. On the other hand, such a representation is always possible in locally convex spaces (cf.§20,7.(5)). We now obtain some more results about cones. (5) Every supporting hyperplane of an algebraically closed cone passes through the vertex of the cone. Let o be the vertex of the cone K. If ux = y is a supporting hyperplane through x0, and if K is contained in the half-space ux^y, then we must have uo = 0^y. If y were negative u(px0) = py would then be less than y for p>l, which is impossible. But it follows from y = 0 that the hyperplane is given by ux = 0, and so it contains o. (6) // an algebraically closed cone lies in one of the half-spaces defined by a hyperplane H, the hyperplane parallel to H through the vertex of the cone is a supporting hyperplane. Again let o be the vertex of K, and let ux^y for all xeK. It follows that y^O. We show that ux^O for all xeK. If ux0 were negative for some x0eK, u(px0) would be less than y for a suitably chosen p>0, and this is not possible. We now give an example of a closed convex cone in the real space //;, p^ 1, for which not every topological boundary point is a point of support of a closed hyperplane. Let C be the set of all x = ({J e lp with t„ ^ 0, n = 1,2,.... C is a closed convex cone with vertex o. Since all the closed hyperplanes of support pass 00 J j through o, by (5), they must be of the form ux= £ v„£„ = 0t where uelp,- + -= 1, «=i P Q or we/00 if p=l. If C lies in wx^O, all the vt must be non-negative. Consequently only those points x which have at least one zero coordinate can belong to closed supporting hyperplanes. But it is easy to see that C has no topologically interior points. It follows that no closed supporting hyperplane passes through any of the points of C for which all the coordinates are non-zero, and these are boundary points.
196 § 17. The separation of convex sets. The Hahn-Banach theorem 6. The Hahn-Banach theorem for normed spaces. Adjoint mappings. We shall now derive certain consequences from the Hahn-Banach theorem: in the next chapter we shall establish these results in a somewhat more general form, for arbitrary locally convex spaces; meanwhile they will serve to complement the theory of normed spaces developed in § 14. In § 14 we evaluated some specific dual spaces; the Hahn-Banach theorem now enables us to show that every general normed space possesses continuous linear functional which do not vanish identically. (1) Given a linear functional l(z) on a linear subspace of a normed space E which satisfies an inequality |/(z)|^M||z||, / can be extended to a continuous linear funtional u, defined on the whole of £, which satisfies \ux\^M\\x\\. If x0 is an element of E, there exists a functional u0eE' with u0x0 = \\x0\\ <md IK|| = 1. The first part comes from 3.(5), and the second part follows from 3.(8), and from the definition of the norm in E' given in § 14, 5.(3). The next statement, which is somewhat sharper, is frequently used: (2) // H is a linear subspace of the normed space E and if x0 is an element of E at distance d from H, then there exists u0eE' with u0x0=l, ||w0|| = l/d and uoz = 0 for all zeH. Proof. By hypothesis ||x0 + y||^d for all y in the closure H of H. + // by the formula l(ax0 + y) = a, ^|a|-d, so that |/(ax0 + y)| = |a| If we define a linear functional on [x0] y then we have ||a^c0 + j^|| = |a| bc0 H— 1 " a|1 ^- Haxo + yll. Using (1), / can be extended to a continuous linear func- d i tional u0 on the whole of E which satisfies |w0x|^-||x||. This shows i d that ||w0||^-, and clearly u0x0 = l(x0)=l and uoz = 0 for all zeH. d On the other hand, if yn is a sequence in H with ||x0 + .yj->d, then l=M0(x0 + >;fi)^||M0|| llxo + ^H, by §14,5.(4), and so KH^-, i.e. 1 a \K\\ = -d- (3) The bidual E" of a normed space E contains £, and on E the norm of the bidual space coincides with the norm of £, i. e. we have (4) ||x||= sup |mx|, for xeE. Proof. Each x0e£ defines a continuous linear functional x0(u) = ux0 on £', and we have |x0(m)| = |m(x0)|^||m||||x0||, so that sup |wx0|
7. The dual space of C(I) 197 ^||x0||. But on the other hand there exists a u0 with ||w0|| = l and w0x0 = ||x0||, by the second part of (1), and so (4) is established. Thus equation (4), which had previously been established for some of the examples of § 14, is now proved in complete generality. In § 14, 6. we also introduced a norm on the space of continuous linear mappings, and in this way made the space 2(E,F) of continuous linear mappings from a normed space E into a normed space F into a normed space. To each mapping Ae2(EF) there corresponds the adjoint mapping A' which maps the algebraic dual space F* into F* (cf.§9,4.). (5) // A is a continuous linear mapping from the normed space E into the normed space F, the adjoint mapping A' maps F' continuously into E\ and \\A'\\ = \\A\\. Proof. A' is defined by the relation (A'v)x = v(Ax), where xeE and veF*. We restrict attention to those v which lie in F'. It follows from § 14, 5.(4), and §14, 6.(4) that \(A>v)x\ = \v(Ax)\^\\v\\\\Ax\\^\\v\\\\A\\\\xl But this means that A'v is bounded by ||j4|| ||i;||, so that A'v is a continuous linear functional on F, and A' maps F' linearly and continuously into E'. Further, the relation ||v4'||^||,4|| follows from \\A'v\\= sup U'v)x\^\\A\\\\v\\. 11*11 ^i Given e>0, there exists an x0eE with ||x0||^l and ||/lx0||>||y4||—c. By (1), corresponding to y0 = Ax0 there exists a v0eF' with ||i;0|| = l and v0y0 = \\y0\\>\\A\\-e. But then \\A'v0\\^:(A'v0)x0 = v0y0>\\A\\-£9 so that \\A'\\>\\A\\-s; i.e. M'|| = ||>1||. (6) // A is a continuous linear mapping from the normed space E into the normed space F, A" is an extension of the continuous linear mapping A to a continuous linear mapping from E" into F". ||j4"|| = ||j4||. A"e2(E\F") and ||/T|| = ||/1||, by (5). By (3), E" => F and F" ^ F. We establish the value of A"x, for xeE. The relation A"x = Ax holds, provided that the two terms represent the same linear functionals on F', and so provided that v(A"x) = v(Ax) for each veF\ But this follows from a double application of the definition of the adjoint maPPmg: v(A"x) = {A'v)x = v(Ax). 7. The dual space of C(/). As a further example of an application of the Hahn-Banach theorem, we shall prove the theorem of F. Riesz characterizing the dual of the (B)-space C(I) of all continuous real- or
198 § 17. The separation of convex sets. The Hahn-Banach theorem complex-valued functions on /= [0,1], i. e. characterizing the space of measures on /. Let u(f) be a continuous linear functional on C(I) which does not vanish identically. If ||w|| = m, then \u(f)\ ^m||/||, where 11/11- sup |/(t)| is the norm of / in C(I). By §14,11., C(I) is a closed linear subspace of L°°(7), and the norm of L°°(7) coincides on C(I) with the norm of C(7). By the Hahn- Banach theorem, u can be extended to a continuous linear functional u defined on the whole of L°°(7), and with the same bound m. In particular the value of u is defined on the function (pceU°(I) which is equal to 0 for t>c and is equal to 1 for t^c, where O^c^l. We set u((pc) = g(c). We show that the function g(c) defined in [0,1] in this way is a function of bounded variation. Suppose that 0 = co<cl <--<cn=l is a decomposition of [0,1] into finitely many subintervals, and let £t = lz^— when 0(c,-)-0(Ci_i) Q(ct) — #(ci-i) + 0> and otherwise let £,- = 0. Then n n / n \ (l) I \g(Ci)-g(Ci-i)\= I sMci)-g(ci-1)) = u( £ ^{(pCi-(pCl_x)) i = \ i = l \i = l J for the function Ysei((Pcl-<Pcl-l) nas absolute value ^1 in [0,1]. It i follows from this that the total variation \J g of g, which is the least " 0 upper bound of the values of £ \g{cv) — #(<Vi)| for all decompositions of [0,1] into finitely many subintervals, is less than or equal to m. The value of u on an arbitrary feC(I) can be expressed in terms of g(c) in the following way: if we divide [0,1] into n equal parts and form the step-function fn(t)= £ f\~)\(Pk(t)~(Pk-i(t))> tnen tne functions k=1 \ / \ n n ' fn converge uniformly to/ and so converge to j Consequently u(f) = \imu(fn)= lim L f\ — \ fn converge uniformly to/ and so converge to/in the sense of the norm. 'k\ fk-\\, .But g*n rl this limit is identical with the Stieltjes integral j f{t)dg, which exists o since/is continuous and g is of bounded variation (cf. Natanson [1], for example, for the results used here about functions of bounded variation and the Stieltjes integral).
7. The dual space of C(/) 199 From the inequality (2) Hf)\ = \f(t)dg ^supLAO|V(flr) = ||/||V(0) 0 0 it follows that ||w||^V^; on the other hand, we showed above that i V g^m = \\u\\. Consequently ||w||=V(g), where g is the function of o bounded variation corresponding to u. Conversely, it follows from (2) that every real or complex function g of bounded variation defines a continuous linear functional on C(7), whose norm is less than or equal to V(g). Which functions g of bounded variation determine a continuous linear functional on C(I) which vanishes identically? (3) \fdg = 0 for all feC(I) if and only if g is equal to some fixed o constant K at 0,1 and all its point of continuity. It is well-known (cf. Natanson [1] p. 219) that the set of points of discontinuity of g is at most countable, and that for each point of discontinuity c0 the limits g(co + 0) from the right and g(c0 — 0) from the left both exist. If g satisfies the given conditions, g is therefore equal to K at 0,1 and on a dense set. Choosing the points of dissection for the approximating sums of the Stieltjes integral from this set, we clearly obtain the value 0 for the integral of any function/ Conversely, suppose that g does not satisfy the conditions. If g(0) =h g(l), i then, putting /= 1, the integral § dg = g(l) — g(0)^0. We can therefore o suppose that g(0) = g(\) = 0. Let c0 be a point of continuity of g at which g(co)=f=0. In a sufficiently small intervall [c0, c0 + e] the variation of g(c) is less than |g(c0)|. This is a consequence of the theorem (cf. X Natanson [1], p. 223) which states that the variation V(g) is continuous b c b at a point of continuity of g, and the relation V(g) = V(g) + Y(g) for a a c 1 0^a<c<b^l (Natanson [1], p. 217). If we consider \ fdg for a o function/which is equal to 1 from 0 to c0, which decreases linearly from 1 to 0 in the interval [c0,c0 + c], and which is equal to 0 from c0 + etol, then the integral from 0 to c0 takes the value g(c0), the integral from
200 § 17. The separation of convex sets. The Hahn-Banach theorem c0to c0 + s takes a value with modulus less than |g(c0)|, and the integral from c0 + e to 1 is equal to 0. Combining these, jfdg^O. This completes the proof of (3). ° A function h(c) of bounded variation on [0,1] is said to be normalized if A(0) = 0 and h(c) = h(c + 0) for 0<c<l. If g(c) is an arbitrary function of bounded variation, the corresponding normalized function g*(c) is obtained by setting g*(0) = 09g*(l) = g(l) i i -g(0) and g*(c) = g(c + 0)-g(c), for 0<c<l. $fdg=$fdg*, since o o g(c) — 0*(c) is equal to 0(0) at 0,1, and all points of continuity of g. Thus every continuous linear functional on C(I) is determined by a normalized g. Conversely every normalized g which does not vanish identically determines a linear functional which does not vanish identically, since a normalized function only satisfies the condition of (3) if it vanishes identically: it is equal to 0 at 0,1 and each point of continuity, and at points of discontinuity it is equal to g(c + 0), which being the limit of values of g at points of continuity must be equal to 0. Finally we shall show that if h(c) is a normalized function the norm i of the continuous linear functional which it defines is equal to V (h). o An arbitrary function g{c) of bounded variation with g(0) = 0 differs from the corresponding normalized function g*(c) only at points of n discontinuity. If we form the sum £l#(c;) — #(c;-i)l onty at points of i continuity, then the least upper bound V of these sums is the same for g and g*. But because of the definition of g* at the points of discontinuity, i V is equal to V (g*)\ consequently the normalized function has the smallest variation of all those functions which define the same normalized function. As we have already shown, every continuous linear functional is determined by a function g of bounded variation whose variation is equal to ||w||, and so the corresponding normalized function g* can only have a variation ^ ||w||; it then follows from (2) that V(gf*) = ||w||. We gather these results together to give the theorem of Riesz [1]: (4) Every continuous linear functional u on the real (respectively complex) (B)-space C(I) can be represented by a real (respectively complex) normalized function hu of bounded variation on [0,1]: (5) u(f)=\fdhu, feC(I). 0
7. The dual space of C{I) 201 The correspondence u-+hu is a norm isomorphism of the space C(I)' of measures on I onto the space V(I) of normalized functions of bounded variation, equipped with the norm (6) \\h\\=V{h). 0 We can now answer the question of whether or not C(I) is reflexive (cf. Riesz-Sz. Nagy[1]): (7) C(I) is not reflexive. It is sufficient to produce a continuous linear functional v0(h) on i V(I) which cannot be represented in the form v0(h)=\ f0dh, f0eC(I). o Now every h has only countably many points of discontinuity, and the 00 sum of the jumps v0(h)= £ (/Kc;)~~ h(ci — 0)), 0<cf^l, clearly satis- i i=i fies \v0(h)\^Y(h). Further v0(h) is linear, so that it is a continuous o linear functional on V(I). If v0(h) were representable by a function f0, then in particular for the function h = \l/c, where ^c = l for x<c and ^c(x) = 0 for x^c, we would have v0(\l/c)= — 1 = — f0{c\ so that f0(c) would be identically equal to 1. But for continuous functions h, i §dh = h(\) — h(0) is in general different from zero whereas v0{h) is o always equal to 0.
CHAPTER FOUR Locally Convex Spaces. Fundamentals The first two paragraphs are concerned with methods of producing new locally convex spaces from given ones. Thus subspaces and quotient spaces of locally convex spaces are again locally convex. The same holds for topological products and locally convex direct sums. The completion of a locally convex space is obtained in a simple way by embedding the space in a topological product of Banach spaces. In § 19, locally convex hulls and locally convex kernels of locally convex spaces are introduced, together with the related ideas of topological inductive and projective limits. The precise distinction between hulls and inductive limits, and likewise between kernels and projective limits, seems to be advisable when giving a systematic account. Some properties of (LF)-spaces are also dealt with in § 19. § 20 begins by examining the dual E of a locally convex space E; the dual pair <£',£> and the weak topology are introduced. The theory of duality for closed linear subspaces, complementary decompositions and linear mappings follows in a simple way. The duality of absolutely convex closed subsets leads to the important properties of polarity. The paragraph ends with the Alaoglu-Bourbaki theorem on the weak compactness of the polar of a neighbourhood of o, and the theorems of Banach-Mackey and of Mackey on the identity of weakly and strongly bounded sets. § 21 deals with the different topologies of a locally convex space, and above all with the strong topology, the Mackey topology and the topology of precompact convergence. A detailed knowledge of these is indispensible, since the coincidence of two of these topologies leads to important structural properties of the space. In particular the topologies of metrizable locally convex spaces are investigated. § 22 applies the theory of duality to the spaces considered in § 18 and § 19. For example we investigate how far it is possible to express the different topologies of a topological product in terms of the corresponding topologies on the factors. We show that these questions do not by any means always have a simple answer in the completely general case. § 18. The definition and simplest properties of locally convex spaces 1. Definition by neighbourhoods, and by semi-norms. A topological vector space E[X] over K (which once again is the field of real or complex numbers) is said to be locally convex if it has base of neighbourhoods U = {Ua} of o consisting of convex sets Ua.
1. Definition by neighbourhoods, and by semi-norms 203 It follows from § 15,1.(3) and § 16,1.(2) that it then also has a base of neighbourhoods of o consisting of absolutely convex sets Ua9 since the convex cover of a circled neighbourhood of o contained in Ua is absolutely convex. Corresponding to § 15, 2., we have (1) Let U = {Ua}, aeA, be a filter-base of absolutely convex absorbent subsets Ua of a real or complex vector space E, with f] Ua = o. If aeA each set pUa, p>0, belongs to U whenever Ua does, then a locally convex space E[%] is defined by taking U as base of neighbourhoods ofo; every locally convex space can arise in this way. This is a simple consequence of § 15, 2.(2), for jUa + jUa = Ua, since Ua is absolutely convex, and so (LT1) is satisfied. In applications, the topology on E is often defined in a more appropriate way by means of a system of semi-norms. We saw in § 16, 4. that every absolutely convex absorbent set has a semi-norm as Minkowski functional, and that the algebraic hull Ua is given by p(x)^\, and the algebraic kernel U{ by p(x)<\. Conversely, starting from a semi-norm p(x) on £, the sets p(x)^\ and p(x)<\ are absolutely convex, absorbent, and algebraically closed and open, respectively. By §16,4., to the absolutely convex neighbourhoods of a locally convex space there correspond continuous semi-norms; conversely, if p(x) is a continuous semi-norm the sets p(x)<£ and p(x)^e are open (respectively closed) ^-neighbourhoods ofo (§ 16,4.(7)). (2) Every semi-norm defined on E[%~\ which is continuous at o is uniformly continuous on the whole of £[£]. This follows directly from the inequality §14,1.(1), which is also valid for semi-norms. (3) Let {Px{x)} be a system of semi-norms on a vector space £, with the property that for each x0 4= o there is at least one pa with pa(x0) 4= 0. // we denote by Ua the set of all xeE for which pa(x)<l, then the sys- n tern of scalar multiples pU, p > 0, of finite intersections U = f] Uai of i= 1 the U^ forms a base of neighbourhoods VLofo for a locally convex topology X on E; this base consists of absolutely convex open sets. Every locally convex space can arise in this way. n The set pU = p f] Uai is given by the inequality p(x)<p, where p(x) i= 1 is the semi-norm sup pa.(x). i = 1,..., n l
204 § 18. The definition and simplest properties of locally convex spaces // instead we take the sets defined by pa(x)^ 1, then in a corresponding way we obtain a base of closed absolutely convex neighbourhoods of o. Proof. The sets p U are absorbent and absolutely convex in £, by the preceding remarks, and they form a filter-base on E. Since for each non-zero x0 there exists a pa with pa(x0) 4= 0, the intersection of all the sets p U is equal to o, and so by (1), E[X] is locally convex. By the preceding remarks the sets p U are open. Conversely, a locally convex space given by (1) can also be given by a system of semi-norms. If the functions pa(x) are the semi-norms corresponding to the sets Ua, the open kernels and closed hulls of the sets p Ua are given by pa(x) < p and pa(x) ^ p respectively, and in either case these sets form a base of neighbourhoods ofo for X. n Finally, the open kernel of the intersection f] Ua.9 which is again absorbent and absolutely convex, is given by the inequality p(x) = sup pai(x) < 1; p(x) is therefore a semi-norm as well. i= l,...,n Let us remark that if {pa} is a system of semi-norms defining the topology of £[£], then the neighbourhoods pa{x)<p will in general only form a sub-base for the filter of neighbourhoods ofo. (4) Let % and %' be the locally convex topologies defined by two systems {pa(x)} and {qp(x)} of semi-norms on E. X is finer than X' if and only if given qp, there exist finitely many pa.y i= 1,..., n, and a p>0 for which the inequality pqp(x)^ sup pai{x) i=l,...,n holds on E. Proof. If Ur and U2 are two algebraically open absorbent absolutely convex sets, the relation L^ <= U2 is equivalent to the inequality p2(x) ^pl(x) between the corresponding Minkowski functional px and p2. % is finer than X' if and only if every ^'-neighbourhood of o of the form qp(x)<l contains a ^-neighbourhood ofo. By (3), this can be 1 taken in the form sup pa.(x)<p, and it therefore has — sup pai(x) i = 1,..., n p i = 1,..., n as Minkowski functional. (4) follows from this. A criterion for the equality of two topologies, each of which is given by a system of semi-norms, follows directly from (4); this corresponds to the Hausdorff criterion for equivalence (§ 2,4.(1)). 2. Metrizable locally convex spaces and (F)-spaces. Every normed space is locally convex, since its topology is defined as in 1.(3) by a single semi-norm, which is in fact a norm.
2. Metrizable locally convex spaces and (F)-spaces 205 Earlier we saw that by no means every metrizable topological vector space is locally convex; indeed the spaces LP, 0<p< 1, which are metrizable, by §15,11., actually have no convex neighbourhood of o other than U itself (§17,11.). A complete metrizable locally convex space is called an (F)-space, after Frechet (cf. the remark of § 15, 11.). Every (B)-space is an (F)-space. Since a metrizable locally convex space has a countable base Un, n= 1,2,..., of neighbourhoods of o, and since each Un contains an absolutely convex neighbourhood of o, the sets Un can all be assumed to be absolutely convex. The sets Uu Ulr\U2, U1nU2nU3,... also form a base of neighbourhoods of o, if Ul9 U2, U3,.'.., do. We therefore obtain (1) The topology of a metrizable locally convex space E[X^\ can always be given by a decreasing sequence U1^> U2^> '" of absolutely 00 convex neighbourhoods of o, with Q Un = o. n= 1 The semi-norms pn(x) corresponding to the neighbourhoods Un then form an increasing sequence p1(x)^p2W = Conversely, starting from a countable system {#;(x)} of semi- norms on a vector space £, the increasing sequence of semi-norms Pi(x)= sup qk(x) defines the same topology, by 1.(3). The sets Un fc=l,...,i consisting of all those xeE with pn(x) < — then form a decreasing n fundamental sequence of neighbourhoods ofo. We therefore have (2) A locally convex space E[X] is metrizable if and only if X can be defined by a countable system of semi-norms. If this is the case X can also be defined by an increasing sequence p1(x)^p2(x)^ ••• of semi-norms. The neighbourhoods pn(x) < —, n= 1,2,..., form a base of neighbourhoods of o for X. n The construction of an (F)-norm on a metrizable topological vector space given in §15,11. depends upon the complicated methods of §6,7.(1). In the locally convex case the following simpler construction is possible: (3) Suppose that E[%~\ is locally convex and metrizable, and that IMIi = IMl2= '" is an increasing sequence of semi-norms which defines X. X is also defined by the (F)-norm (4) ||x|| = Y - -!!^- ; .= 12" l + IML'
206 § 18. The definition and simplest properties of locally convex spaces X is therefore defined by a translation-invariant metric given by \x,y\ = \\x-y\\. Proof. First we establish the (F)-norm properties. (Fl) is trivial, and (F2) is satisfied, since for each non-zero x0 there is a semi-norm ||x|L with HxolL + 0. a b If 0<a^b, ^ . \+a \+b It follows from ||Ax||ii= \X\ \\x\\n^ \\x\\n9 for |A|^1, that ||Ax||g||x|| for |/|^ 1, so that (F3) is satisfied. ||x + y\\ ^ ||x|| + ||y|| follows directly from llx+ylL ^ HxlL+llylL ^ IML + ||y|L 1 + \\x + y\\H ~ 1 + ||x|L+ \\y\\n ~ 1 + ||x|L 1 + \\y\\n so that (F4) is satisfied. Next we show that the system of ^-neighbourhoods ofo is equivalent to the metric defined on E by (4). 1 a) The neighbourhood ||x|| < -^, /c^l, contains the I-neighbour- hood ||x||k+1 <^TT 1+|, <||x||n. If x satisfies ||x||fc+1 <^rrT, then HxH^--- = 1 1 k + i 1 a 1 1 ^ W/t+ i < -ttt as well, so that ||x|| < -r— Y h Y — < —. b) The ^-neighbourhood ||x||m < — contains the neighbourhood 1 2 llxll < yn+k+l ' If x satisfies llxll < —-ttt, then -^— < —-—r, so that ii ii 2m+k + 1 2m l + ||x||m 2m+k+1 x|L 1 / 1 \ 1 < -ttt. It follows from this that llxIL 1 - -7—7 < l + ||x|L ' 2k+1' u"umY 2k + l) 2fc+1' and so that ||x||m <^. Since the two neighbourhood systems are equivalent, it also follows that (F5) and (F6) are satisfied, since the corresponding assertions are certainly true for I-convergence. We observe that for this metric the distance between two points is always less than or equal to one. 3. Subspaces, quotient spaces and topological products of locally convex spaces. The results obtained in §15,4. can be extended, in the locally convex case.
3. Subspaces, quotient spaces and topological products of locally convex spaces 207 (1) Every linear subspace H of a locally convex space £[3Q is again locally convex, in the induced topology. For U n H is absolutely convex if H is. (2) Every quotient space E/H of a locally convex space E[Z~] by a closed linear subspace H is locally convex in the induced topology X. If the topology X of £[£] is given by a system {p^x)} of semi- norms which always contains a py^sup(pa,pp) when it contains pa and pfi9 then the induced topology of E/H is given by the system {pa(x)}, where pa(x) is the semi-norm (3) pa(x)=infpa(x), xex and x is a coset in E/H. Proof. K(U) is absolutely convex if U is, where K is the canonical mapping of E onto E/H. The quotient space topology therefore has a base of absolutely convex neighbourhoods of o. The assertion about semi-norms is proved in the same way as in §14,4.(1). It follows from (2) and § 15,11.(4) that (4) Every quotient space of an (F)-space by a closed linear subspace is again an (F)-space. If £[3T| is an arbitrary complete locally convex space, E/H need not be complete (cf. § 23, 5. and § 31, 6.). Similarly for topological products we have (5) The topological product £[3T| = TT JEa[Xj of locally convex spaces is again locally convex. For the neighbourhoods TT Wa, where Wa is an absolutely convex a neighbourhood l/a of o for finitely many a, and W0i=E0i otherwise, are clearly absolutely convex. If, for fixed a, {ppixj} is a system of semi-norms on Ea which defines the topology %a, and if we denote by pp(x) the seminorm on E defined by P°p(x)=P°p(xa)> where x = (xa)eE, then the system of all the semi- norms p°p(x) defines the product topology on E. Let us recall that it was shown in § 5, 7. that the completion of a topological product is equal to the topological product of the completions. It follows from this and from the ideas which have just been used that (6) The topological product of metrizable locally convex spaces is metrizable if and only if the product has finitely or countably many factors.
208 § 18. The definition and simplest properties of locally convex spaces The topological product of countably many (F)-spaces is again an (F)-space. The next result gives a certain way of looking at all possible locally convex spaces (cf. § 10, 7.(9) as well); (7) Every locally convex space F[3T] is topologically isomorphic to a linear subspace E of a topological product of (B)-spaces. E is complete if and only if E is closed. Proof. Let {pa(x)} be a system of semi-norms on E which defines the topology X. If Na is the null-space of pa(x), then, by § 14,1. (4), F/Na = Ea is a normed space under pa(x(?) = pa(x), where xa is the coset of x in Fa. We form the completion of Fa of Fa (§ 15,3.). Let F be the topological product TTFa of the (B)-spaces Ea. The extension of the norm pa to Fa a is again denoted by pa. We now map each xeE to the corresponding element x = (xa) ofF. This mapping is one-one and linear, and it maps E onto a linear sub- space E of F. We saw above that the topology of F is defined by the semi-norms p0i(x) = p0i(x(X). Since pa(x) = pa(xa) = pa(x), corresponding semi-norms pa and pa take the same value on corresponding elements xeE and xeF, and so the correspondence is a topological isomorphism. It follows from the completeness of £ that E is closed; conversely, if E is closed in F, E is complete, since F is, and so therefore is E[X~\. (8) Every real locally convex space E[X~] can be embedded in a complex locally convex space F[2/] in such a way that F = E@iE and the topologies induced on E and i E by X' coincide with X. Proof. We set F[2/] equal to the topological product ExE, and define multiplication by i by the equation i(x,y) = ( — y,x). In this way F becomes a complex vector space. We identify E with the real subspace of all (x,o). Then (x,y) = x + iy, so that iE is the real subspace consisting of all points (0,3;), and F = E@iE. By definition, the product topology X' induces the topology X on E and on iE. If £ is a real normed space with norm p(x), then by §14,4.(5) the expressions sup(p(x), p{y)\ p(x) + p(y), |/p2(x) + p2(y\ etc., are norms on F whose restrictions to E and to iE coincide with the given norm. 4. The completion of a locally convex space. By §15,3.(1), every topological vector space has a completion. (1) The completion F[2T| of a locally convex space E[X^ is locally convex. For by § 15, 3.(1) we obtain a base of neighbourhoods of o in E by forming the closures Ua in FpX] of a base of absolutely convex neigh-
4. The completion of a locally convex space 209 bourhoods Ua of o in E. But these are again absolutely convex, by §16,1.(5). (2) // the topology of E[X] is defined by a system {pa} of semi-norms, the topology of £[2] is defined by the system {pa}, where pa is the uniquely defined continuous extension of pa(x) to E. Proof. Every semi-norm pa{x) is uniformly continuous on £, by 1.(2), and so it has a uniformly continuous extension pjy) to £, by § 5,4.(4). Properties (N1), (N3) and (N4) of § 14,1. are still satisfied, so that pa(y) is a continuous semi-norm on E. We denote the neighbourhood pa(x)<c of o in E by l/a£, and its closure in E by UaE. We denote the neighbourhood pa{y)<& of o in E by Ka£; it is open, since pa is continuous. Its closure in E we denote by Pa£; it is given by p^(y)^s, by § 16, 4. To prove that X is defined by the system {pa} it is sufficient to show that CaE=VaE. Because pa is continuous, we clearly have pa{z)^& for each zeUaE, so that UaE a VaE. Since a point z in the open set VaE is a closure point of UaE= VaenE we have VaE cz (7a£, so that Fa£ cz Uae. The construction of the completion of a topological vector space given in § 15, 3. depends upon the construction of the completion of a uniform space given in § 5, 5., and this is not particularly simple. Using 3.(7), we can give a rather simpler construction which only depends upon the construction of the completion of a normed space given in §14,3. A further construction will be given in § 21,9. We recapitulate the assertion: (3) Every locally convex space E[^t] can be embedded in a smallest complete locally convex space £[2T|, which is unique up to topological isomorphism. The closures in E of the members of a base of neighbourhoods of o in E form a base of neighbourhoods of o in E. Proof. By 3.(7), we can embed E\1L\ in a topological product F of (B)-spaces. Because F is complete, the closure E of E in F is a complete locally convex space; this establishes the existence of a completion £[X]. Next we show that the closures Ua in £ of a base of open neighbourhoods Ua of E form a base of neighbourhoods of o in E. Since H is the topology induced on E by I, for each Ua there is an open neighbourhood Ua of o in £, with Ua=U0inE. Since Ua is open and E is dense in Ua9 Ua is dense in Ua. Hence Ua cz Ua, Ua cz Ua, and so £?a= Ua is a neighbourhood of o in E. If now V is an arbitrary closed neighbourhood of o in £, V= Kn £ is a neighbourhood of_o in E. Thus, for a suitable l/a, Ua cz V, and so C7a cz V cz V. The sets Ua therefore form a base of neighbourhoods of o in E[X~]. 14 Kothe, Topological Vector Spaces I
210 § 18. The definition and simplest properties of locally convex spaces Two different completions of E can be mapped in a one-one way onto each other in a way which makes the limits of the same Cauchy nets correspond to each other. Under this the closures of the sets Ua correspond, and so the two completions of E are topologically isomorphic, since the correspondence also preserves the linear operations. In many situations the locally convex spaces under consideration are not complete. The two following weaker properties are frequently useful: a locally convex space E[X~] is said to be sequentially complete if every X-Cauchy sequence has a limit in E[Z~\, and it is said to be quasi-complete if every bounded closed set of E[I] is complete. Clearly every complete space is quasi-complete, and every quasi- complete space is sequentially complete. Suppose that a second finer locally convex topology X' is given on a locally convex space E\X~\. What relation is there between the completions JEpT] and £[£]? The identity mapping / from £[3/] into E[X~] is continuous, and so by § 5,4. (4) there is a uniquely determined continuous extension I which maps £[£'] into E[%~\. In general this linear mapping need not still be one-one, so that it is only the quotient of £[27] by /"^o) which is embedded in E[2T|. For example, let £[£] be a (B)-space and let X be the topology defined jon E by a finer norm which is not equivalent to the X-norm. If / were one-one, E[X'~\ and E[X] would be topologically isomorphic, by the Banach-Schauder theorem [cf. § 15,12.(2)], and this is impossible. However the following important completeness criterion holds, even for arbitrary topological vector spaces: (4) Suppose that a second finer topology X' is given on the topological vector space E[_X~\. Suppose that X' has a base of neighbourhoods of o consisting of X-closed sets. Under these hypotheses a) // a X'-Cauchy filter converges to x0 as a X-Cauchy filter on £, then it also converges to x0 as a X'-Cauchy filter. b) Every subset of E which is complete (respectively sequentially complete) with respect to X is also complete (respectively sequentially complete) with respect to X'. c) The identity mapping I from E\%'~\ onto E[%\ can be uniquely extended to a continuous embedding I of £[£'] into E\Xi\. Proofofa). A X'-Cauchy filter g is certainly a X-Cauchy filter, since X is coarser than X'. Suppose that x0 is its X-limit. To each X-closed circled ^'neighbourhood U of o in E there is an Faeg which is small of order U. If yeFa, F*ay+U. Since y+U is I-closed, x0
5. The locally convex direct sum of locally convex spaces 211 lies in y+U, as it is a X-closure point of F*. It then follows from F* <= x0 + U + U that x0 is also I'-limit of ft. b) follows directly from a) and from the corresponding assertion for sequences, which is proved in an analogous way. Proof of c). We must show that / is one-one. Let z be an element of £[£'] with 7z=o. There exists a 3/-Cauchy filter ft={Fa] in E with limft = z. The sets Fa form the base of a J'-Cauchy filter ft' in £[2/] whose limit is again z. Under the continuous mapping 7, ft' is mapped into a X-Cauchy filter ft = 7(ft') whose limit 7z is o, by hypothesis. The restriction ftn£ of ft to E is coarser than ft: for if I(M) is an arbitrary set of the filter ft, where Me ft', and if Fa c M, then Fa = 7(Fa) c 7(M), so that Fa <= 7(M) n £. Thus ft n E is a filter on £, and as it is the restriction of ft it is a 2-Cauchy filter with limit o. But then the finer J-Cauchy filter ft has £-limit o. It now follows from a) that ft, considered as a £'-Cauchy filter, also has o as limit, so that z=o. (4) is due to N. Bourbaki [6] and W. Robertson [2]. The latter work contains a detailed investigation of this and related problems. 5. The locally convex direct sum of locally convex spaces. In §7,8. we introduced the direct sum E= © Ea of vector spaces Ea. This is the a subspace of T\Ea consisting of only those elements x = (xa), xae£a, a which have finitely many non-zero xa. We denote the embedding of Ea into E by 7a. This is the mapping which sends the element x0LeE0i to the element xeE whose a-th coordinate is equal to xa, and all of whose other coordinates vanish. The locally convex direct sum E[Z~] = ©£a[Ja] of the loot cally convex spaces £a[Xa] is defined to be the direct sum E of the spaces Ea equipped with the finest locally convex topology for which each of the embeddings Ia of Ea into E is continuous. It is easy to give a base of neighbourhoods of o for E\Z~\. (1) If for each fixed a, {Up} forms a base of neighbourhoods of 0 in £a, then the absolutely convex covers \~ IJJJp) form a base of neighbourhoods of o in ©£a[Ia]. a a For if U is an absolutely convex neighbourhood of o for some topology on E for which la is continuous, then U must contain a set Ia(Up). This holds for each a; but then \~ Ia(Up) also lies in £/, since U is abet solutely convex. The finest of all these locally convex topologies on E is however clearly the one for which all such sets [~ I^p) f°rm a base of neighbourhoods ofo. In what follows we shall write Up instead of Ia{Up), provided that there is no danger of misunderstanding. 14"
212 § 18. The definition and simplest properties of locally convex spaces (2) The topology X of © Ea[Xa~] is Hausdorff, so that ©£a|X] a a is again a locally convex space. If /? runs through a subset of the indices a, © Ep [Xp] is a closed subspace of ©£a[2a], and the topology induced on ©Ep[Xp] by X is a p the locally convex direct sum topology. In particular X induces the topology Xa on each Ea. n n For finitely many summands, © £a,.pXaJ and TT £ai[IaJ are topological^ isomorphic. I_1 I_1 Proof. The topology induced onH = © Ep[Xp] by X is determined by the neighbourhoods (p Ua)nH = [~ Up of o; these also determine the locally convex direct sum topology on H. n n The neighbourhood [~ Uai of o in © £a.pXa.] lies in the neigh- i = 1 i = 1 bourhood TT Uai of o in TT EjXal; on the other hand TT \-Uttl i=l r=l ' ' i=l \n n lies in \~ U*\ so that the product topology and the sum topology i = 1 n n coincide on © £a. = TT Ea.. i = 1 ' i = 1 n Since TT £a.[IaJ is Hausdorff, and since each element xe© £a[2J lies in some © Ea.\Xa^\, there is a ^-neighbourhood of o separating x i= 1 and o, and so X is Hausdorff. E = H@H', with H'= ® Ey[%y~], where y runs through the indices y a which are different from the /?. By §15,8.(1) this decomposition is topological, since the projections onto H and H' are continuous. // and H', being the null spaces of these projections, are closed. (3) // the spaces £a[2a] are complete locally convex spaces, their locally convex direct sum is also complete; consequently, we have ©£a[2a]^©£a[2a]. a a We shall show that E[X~] = ©£a[IJ is complete, when the spaces £a[Xa] are complete. a Let $={FP} be a 3>Cauchy filter on E. Since the projection Pa of £ onto Ea is continuous, Pa(g) is also a Cauchy filter on Ea. Since £a is complete, Pa(5) has a limit xa. We assert that only finitely many xa can be different from zero. For otherwise there would be a sequence xai+o, i=l,2,.... For each xai there exists a closed absolutely convex ^.-neighbourhood U"1 of o with
5. The locally convex direct sum of locally convex spaces 213 o<Jxai+£/"'. We form a ^-neighbourhood V=Y~V\ with Ka'=C/a\ a f=l,2,... . Let Fp be small of order V. For an element z = (za)eFp we have P0Liz = zOLiePOii(Fp). Pa.{Ffi) is small of order Pai(K)=t/ai. Because Uai is closed and xa. is a closure point of Pai(Fp) we have xa.eza.+ l/ai, or za.exa. + £/a*, since £/a'' is absolutely convex. Consequently za.=j=o for £ = 1,2,..., so that z is an element of £ with infinitely many nonzero components in the spaces £a; this gives a contradiction. Now let x = xai + ••• +xan be the sum of the non-zero xa. We still have to show that x = limg. Let U = [~ Ua be a ^-neighbourhood of o. a Further let Fp be small of order U and let z be an arbitrary element of Ffi. m There is a finite sum H = ® Eai [£a.] in which both x and z lie. If P is i= 1 the projection of E onto //, then by (2) P(U)cz U and P(U) is a neighbourhood of o in H. It follows from Fp az+U that P(F*) c Pz + P(t7) = z + P(l/). Since P(g) has x as limit in tf, xez + 2P(l/), so that zex + 2P(C/)<=x + 2£/ and Fpcz:x + 2U. Since this holds for each U and for each Fp which is small of order [/, x = lim g. (4) Every bounded subset B of £[3T| = ©-EaP^J zs a bounded sub- n a s^t o/ a finite sum © £a,.[£a.], and so is contained in a set of the form n i=l © Pa. cz E where Ba. is bounded in £a,[£aJ. i= 1 Proof. The projection P(X(B) = B0i of B in Ea is bounded in Ea[£a], by (2). We must show that there can be only finitely many at with Bai =j=o. Let us suppose the contrary. It is sufficient to consider the case 00 © £,[2,], with £f=)=o for each i. We can also suppose that for each / i= 1 there is an xii) = {xil\xii\...)eB with x|°#o and x{^=o for k>i. Our assumption will be disproved if we can find a neighbourhood U 1 ,.. of o which contains none of the points — x(0, for then there would be i no scalar multiple of U containing all the x(0, so that B would not be bounded. For each i we determine an absolutely convex neighbourhood Ut 1 of o in Et which does not contain — x\l\ and we put U = \~ £/,. Then i i=i we have Pi(U)=Ui and pi-x(0) = -xj°$t/f. But it follows from this that -x(0$£/. i A vector space E can always be represented as the direct sum of one-dimensional spaces £a, £=©xaK, where {xa} is a basis for E.
214 § 18. The definition and simplest properties of locally convex spaces £, considered as the locally convex direct sum of the spaces xa K, has a topology X, which, by (1), has as base of neighbourhoods of o all absolutely convex absorbent sets, i.e. all absolutely convex a-bodies. This topology is clearly the finest of all the locally convex topologies on E. With the notation introduced in § 7, 5. we therefore have (5) The topology on the locally convex direct sum (pd{K) of d one- dimensional spaces topologically isomorphic to K is the finest possible locally convex topology. An immediate consequence of (4) is (6) Every bounded subset ofqyd(K) is finite-dimensional and relatively compact. The introduction of the topology X of the direct sum given here provides a certain contrast with the procedure in the case of linearly topologized spaces (§ 10, 2.). There we took the sets © Ua as base of neighbourhoods of o. In the a present case we can also proceed in this way. If £a[2J are arbitrary topological vector spaces, we define the topology X' on E = © Ea by taking as base of neighbourhoods of o sets © C/a, where Ua is an arbitrary neighbourhood of o in £apXJ. In the locally convex case we clearly obtain a weaker topology than X, since r~ Ua <= © Ua- We call £[£'] the topological direct sum of the spaces £a[2J. The same results hold for the topological direct sum: (7) £[£'] = (©£a)pX'] is locally convex for locally convex £«[$«], E{X'~\ is complete for complete £apXa], and the bounded sets are contained in the sets © Bai, so that in the locally convex case they coincide with the X-bounded sets. The proof, which is almost word for word the same as for the locally convex direct sum, is left to the reader. (8) In the locally convex case, if there are countably many summands £,[£,], 00 then X and X coincide on © E(; for more than countably many summands, X can be different from X. i=l °° In the countable case X is equal to X since every ^-neighbourhood \~ Ul oo 1 '=1 contains the ^'-neighbourhood © —t Ul. i= 1 2 X and X are different on (pdif d>K0: in each one-dimensional space Ea = xa K let the neighbourhood Ua of o be given by |£J^1. Then \~ Ua consists of all a x = {£a)e(pd with 5Z l^«l = 1- But for every X-neighbourhood of o ©pa[/a there a a exists an e0 for which pa^e0 f°r uncountable many a, so that there are always elements x in ©pa[/a with arbitrarily large £|£J. The reason for giving preference to the topology X rather than to X will become apparent when we consider the theory of duality (§ 22, 5.).
1. The locally convex hull of locally convex spaces 215 § 19. Locally convex hulls and kernels, inductive and projective limits of locally concex spaces 1. The locally convex hull of locally convex spaces. Locally convex direct sums and topological products of locally convex spaces are special cases of more general ideas which we now consider. We begin with the locally convex hull and the topological inductive limit, which are particularly important for applications. If a vector space E is the linear span of certain linear subspaces Ea, we write E = YJEa. Of particular interest for us is the case where each £a a is given as the linear image Aa(Fa) of a vector space Fa. We then write a A special case of such a linear span is the direct sum E=®E0L\ conversely we have (1) Every linear span E = Y^Aa(Fa) is isomorphic to a quotient E = (®F„)/H. Proof. A linear mapping A from © Fa onto E = £/la(Fa) is defined a a by A\Yx0\=Y^^olx^ x<xeF<x- If H ls tne null-space N[A] of A in © Fa, then E is isomorphic to E = (® Fa)/H. Clearly we have the converse: (2) Every quotient E = (©Fj/H is equal to the linear span £ = ^Ka(Fa), where Ka is the restriction to Fa of the canonical homo- a ^ morphism K from © Fa onto E. a If the Fa are locally convex spaces Fa[£a], we can try to introduce as natural a locally convex topology as possible on the linear span E = YJA0i(F0l). By analogy with the special case of the locally convex a direct sum, the finest locally convex topology £ for which all the /la are continuous mappings from Fa into E suggests itself. An absolutely convex subset U of E is then a ^-neighbourhood of o if A{~l)(U) is a ^-neighbourhood U* of o in Fa, for each a. All the sets \~ Aa(U% where each set (7a is taken to be a ^-neighbourhood of o in Fa[IJ, therefore form a base of ^-neighbourhoods ofo in E. This topology H need not however always be Hausdorff, as we shall presently see. But if this is the case, E[X] = ^a(Fa[IJ) is called the a locally convex hull of the ,4a(Fa|jXa]), and Z is called the hull topology on E.
216 §19. Locally convex hulls and kernels We can say more about the algebraic isomorphism (1): (3) Every locally convex hull £[3] = X^M^aKx]) *'s topologically isomorphic to a quotient E = (© FJ/H of the locally convex sum of the Fa[IJ by a closed linear subspace H. Proof. Under the mapping A(YJx^j=YJA0ix(X °f ©^Pa] onto a a E[X], the neighbourhood \~ U* of o in ©Fa[Ia] is mapped onto the a a ^-neighbourhood \~~ A^U") of o in £. ,4 is therefore a topological a homomorphism. By §15,4.(4), A is a topological isomorphism of (®F^j/H onto jE[I], where H is the closed null-space of A. The topology on (®FA/H is therefore the quotient-space topology. The stronger assertion corresponding to (2) is clearly also true: (4) Every quotient E = (© FJ/H of a locally convex direct sum of locally convex spaces F^%^\ by a closed linear subspace H is topologically isomorphic to the locally convex hull XjKa(^a[3a]), where Ka is the a restriction to Fa of the canonical homomorphism K of © Fa onto E. a Consequently it is clear that the hull topology on a linear span ^/4a(Fa[IJ) is not Hausdorff if and only if the null-space H of the a mapping A which establishes the algebraic isomorphism (1) is not closed (cf. §10,7.). Remark. By (4), a quotient space (£/#)[£] of £[£] can be considered as a locally convex hull; we have (£/H)pI] = K(E\_(X]), where K is the canonical mapping from E onto E/H. (5) Suppose on the one hand that E is equal to ^a(Fa[IJ), and on a the other that it is equal to X^G^pXJ?]). Let H be the hull topology P corresponding to the first representation, and let %' be that corresponding to the second. If for each a there is a /? for which Aa(Fa) cz Bp(Gp), and if the topology 3^ defined on Bp(Gp) by Bp induces a coarser topology on Aa(Fa) than the topology 3a defined on Aa(Fa) by Aa, then X is coarser then 3 on E. For if V is an absolutely convex T-neighbourhood of o, VnBp(Gp) is the 1^-image of a ^-neighbourhood of o in Gp, and by hypothesis this contains the /la-image of a ^-neighbourhood of o in Fa. V is therefore also a ^-neighbourhood of o. If the two representations define the same topology on E, we speak of equivalent defining systems.
2. The inductive limit of vector spaces 217 (6) The formation of locally convex hulls is transitive. If E[Z] = YJAJYJBpx(Fp^PA £[!] is also equal to 2^ A;xBp(x(Fp(x[(Xp(x']). This is trivial for the vector space properties, and the fact that the hull topologies coincide follows from \~ AJ\~BB (l/H = r AaBfiu{U'-). X,P<x (7) A /mear mapping B from a locally convex hull E[i] = X^4a(£a|jXa]) a (respectively X£a[£a]) *nt0 a locally convex space FpX'] is continuous a if and only if all the mappings BAa (respectively all the restrictions of B to the spaces EJ are continuous mappings from £a[£j into FpX']. Proof. The condition is clearly necessary. On the other hand, if each BAa (respectively BIa, where Ia is the embedding of Ea into E) is continuous, then given an absolutely convex neighbourhood V of o in F there is always a ^-neighbourhood Ua of o with BA^U^a V. But then B(r AJJJ*)\czV, so that B is continuous. An analogous proof shows that a collection M of linear mappings B is equicontinuous if and only if the mappings BAa are equicontinuous, for each a. 2. The inductive limit of vector spaces. The ideas which we shall consider in this number are of a purely algebraic nature, and they are also valid for vector spaces over arbitrary fields. Let E = YjAziFJ be a linear span. Suppose that the index set A is a a directed set, and that for each pair a</? there is a linear mapping Apa from Fa into Fp for which (1) A0L=ApApa for a</?. Further, we put Aaa equal to the identity mapping from Fa onto itself. If a^jS, we always have Aa(Fa)cz Ap(Fp). (2) //' the mappings Aa, aeA, in (1) are one-one, then so are the mappings Apa, and we have (3) AyfiAfia = Aya for a^/^y. To prove this, multiply both sides of (3) by Ay and apply (1). This situation arises in many cases. For example if E = YJE0L, if the a indices a form a directed set A and if Ea cz Ep for a < /?, then we need only set Aa equal to the embedding 7a of Ea in E and APa equal to the embedding Ip<x of £a in Ep, and (1) is satisfied. In this case (3) is also satisfied.
218 §19. Locally convex hulls and kernels We can even express every linear span £ = £/la(Fa), aeA, in this a way: for every finite subset A of A we form the finite direct sum FA = © F5 <5eA and define the mapping AA from FA into E by setting AL Yjxd = Z^<5X<5- 5 d The set of A's becomes a directed set when we put A^A2 whenever Axcz A2. Finally for AX^A2 we define ALlLl to be the embedding of FAl into FAl. In this way, £ = £/4A(FA), anc* (!) *s satisfied. (3) also holds for the mappings ALlLr A If the situation described at the beginning holds, it is transformed by the isomorphism 1.(1) in the following way: (4) If E = Yj Aa(Fa\ and if mappings Ap<x from Fa into Fp are given a ^ which satisfy (1), then the quotient space E = £ Ka(Fa) isomorphic to E satisfies the following relations: a (5) K, = KpApoL for a<j3. Proof. The mapping A of the proof of 1.(1) determines the one-one mapping A from E = K[@ Fj onto £, and so by 1. (2) £ Aa xa = A K \Y xa) = AYJK<xx0i' The element Aaxa therefore corresponds in a one-one a way to the element Kaxa. It therefore follows from Aaxa = Ap(Ap<xxa) that Kax(X = Kp{Ap(Xx(X) for all xaeFa, which implies (5). So far we have started from a space £, together with subspaces Aa(Fa) and mappings Ap<x from Fa into F^. It is natural to ask how far E is determined by the Fa and the Ap<x, and whether, given Fa and Ap<x, we can find a space £ and mappings Aa which satisfy (1). An answer is given by (6) Suppose that Fa is a directed system of vector spaces, such that for each pair a</? there is a linear mapping Ap<x from Fa into Fp, and such that (7) AypApa = Ayoi for a<j8<y. We denote by H0 the linear span in © Fa of all the elements xa— /Laxa, a with xaeFa and a</?. // H 3 H0 is a linear subspace of © Fa and if K is the canonical homomorphism oj © Fa onto [® Fa)/H, then (® Fa)/H = X Ka(FJ, and (5) holds. a This is an easy consequence of 1. (2). For it follows from H 3 H0 that K{xa — Ap(XxJ = o, so that Kax(X = K(XApax(X for all xaefa, and so (5) is satisfied.
2. The inductive limit of vector spaces 219 Taking H = H0, the uniquely determined space (© Fa)/H0 = £ K[0){Fa) a is called the inductive limit of the spaces Fa with respect to the mappings Ap<l, and it is denoted by lim ApoL(Fa). Even when all the mappings Ap<x are non-zero it can happen that lim Ap<x{FJ consists of the element o alone. So far we have made no use of (7); we now use it to determine H0 exactly. n (8) The element Y xa. of ® Fa lies in H0 if and only if there exists i = i a ]8^af, i= 1,..., w, with n (9) Z 'V*^0' i= 1 n n Proof. If (9) holds, then £ xa. = £ (xa.— /I^xJgHq. Conversely i = 1 i = 1 if y^ft the relation i4yaxa — Ay/j(y4/?axa) = o follows from (7), and so we obtain an equation of the form (9) for xa — Apaxa. By choosing a sufficiently large /?, (9) also holds for a linear combination of elements xai~Api0iixai, and indeed this holds for all /?^/?,-, i= 1,..., w. From (8) we obtain a certain converse to (2): (9) // the mappings Ap<x in lim Apa(Fa) = £ K{°\FJ are all one-one, a then so also are the mappings X(a0). For xaeFa, (9) is equivalent to a relation Apax(X = o. The hypothesis that Apa is one-one means that this is only possible if xa = o. Thus FanH0 = o. In particular it follows from this that if the mappings Apa are one- one, the inductive limit is different from zero, provided that the Fa are different from zero. (11) // E = Y, Aa{Fa), and if Ap<x is a system of mappings satisfying (1) a and (3), then E is always a homomorphic image of lim Apa(F^. A necessary and sufficient condition for E to be isomorphic to lim Apa(Fa) is that an equation A0ix(X = o, xa=}=o, holds in E if and only if there is a /?>a (depending upon xj for which Apaxa = o. Proof. If, using (4), we go over from E to the isomorphic space E = £ Ka{FJ = (0 Fa)/H, then (5) holds, so that Kaxa = KpApaxa. This a means that K(xa — Ap<xxJ = o, and consequently x^ — Ap^x^eH. Thus H 3 H0, and so E is the homomorphic image of (© F^/Hq = lim Ap^FJ.
220 §19. Locally convex hulls and kernels Further Aaxa = o holds in E if and only if Kaxa = o holds in E. By (8), X(a0)xa = o, i.e. xaeH0, if and only if APaxa = o for a suitable /?>a. Thus if it always follows from Kaxa = o that Ap<xx0i = o for a suitable /?>a, then Xa and X(a0) have the same null-space in Fa. We still have to show that the equation H = H0 follows from this. Since H=>H0, it follows from K(£xa.)=o for P^<x-w->*n tnat K(YJAp(Xix(Xi) = o; from this it follows that K,(E ^ *«) = *HI ^V- xat) = K<0)(I *..) = °> i.e. H0=>tf. (11) gives conditions for a linear span to be representable as an inductive limit. A linear span of the special form E = YJE0C, with Ea<^Ep for a</?, a is always equal to lim Ipa{Ea)9 where Ip<x is the embedding of Ea into Ep. We also write lim Ea for this, and speak of the inductive limit of the directed system of spaces Ea. As a special case, a direct sum © Ea is the inductive limit of its finite partial sums. Likewise, by (11) and (2) a linear span E = £ 4a(Fa) with a system a of mappings Ap<x satisfying (1) and (3) can be represented as lim APa(F^ if the mappings Aa are all one-one. 3. The topological inductive limit of locally convex spaces. If the vector spaces Fa of the preceding number are locally convex, then we can proceed as in 1. Let FapXa] be a directed system of locally convex spaces, and let {Apa} be a system of continuous mappings from Fa[IJ into i^pX^], for a</?, for which (1) AyPAPa = Aya for oi<p<y. If H0 is again the linear span of the xa — Aaaxae © Fa[IJ, and if H0 a is closed in © Fa[Ia], then the hull topology (respectively quotient topology) on^m^Fj^X^H^Pa])^^^^)/^ is Haus- dorff. ""* The space lim /^(f^pXJ, equipped with this topology % is called the topological inductive limit of Ap^FJ. By the final remarks of the preceding number, a locally convex hull £[^] = Z£a[^al w*tri Ea^Ep for a<]8 is the topological inductive
3. The topological inductive limit of locally convex spaces 221 limit \imIpa(Ea[Zj) = \imEa[Za] of the spaces £a[3j if and only ifZp induces a weaker topology than 3a on £a, for a</?. 00 In particular, the locally convex hull £ £„[£„] of an increasing n= 1 sequence ^ip^] cz £2[3;2] cz ••• is the topological inductive limit of the spaces £„[£„] if and only if the topology 3„ + 1 induces a weaker topology than 3„ on En, for each n. Every locally convex hull £[2] = £Ea[Ia] witn an arbitrary a index set can be considered as the topological inductive limit of the locally convex hulls of all the finite collections of £a[3a]. (2) Suppose that we have an inductive limit E[X] = lim /4/3a(Fa[^Xa]) of locally convex spaces Fa|jXa], and that T is a cofinal subset of the directed index set A. As y and d run through all pairs with y<d in l~, then E1[X1] = limAdy(Fy[%y]) also exists, and it is topologically isomorphic to £[£]. "-* P r o o f. T is also a directed set. We write E as the linear span £ X(a0)(Fa), and^as £ K(y0)'(Fy). Since it follows from a<y that K[°\Fa) c jK(y0)(Fy), a and since, because T is cofinal, there always exists a y^a for each aeA, E is also equal to the linear span ]T K{y0)(Fy). To each element J| K^Xy.eE we now make correspond the element £ K^'x^eE^^. i=l w i=l By 2.(8) the element £ K^x^ is equal to o if and only if there is a i= 1 n jS^^, f=l,...,w, with Y, Aprxr = o in © Fa. But then if deT is such n that <5^j3, we also have £ /4dy.xy. = o. Therefore it follows from 2.(8) that £ K^ xy. e £ is equal to o if and only if the corresponding element Y K(y®yxyieEl vanishes. The correspondence between E and £x is i=l therefore an algebraic isomorphism. The hull topology X on E is the finest locally convex topology for which all the mappings X(a0), aeA, are continuous. But it is enough to require that all the mappings Ky0) are continuous, since it follows from X(a0) = X(y°Mya for a<y and from the continuity of Aya that X(a0) is continuous. This implies that E[i] and li^[3^] are topologically isomorphic. We now return to the general situation considered at the beginning of the preceding number. Suppose now that £[3Q = £ Aa(Fa[Ia]) is
222 § 19. Locally convex hulls and kernels the locally convex hull of a directed system {Fa[Ia]}, and that to each pair a</? there corresponds a continuous linear mapping Ap<x of Fa into Fp, for which again (3) Aa = AfiAfia for a<j3 and (4) AypAPa = Aya for a </?<?. We saw (cf. 2.(11)) that E is the homomorphic image of lim Ap<x(FJ. It is easy to see that this homomorphism is also a topological homo- morphism provided that the topological inductive limit of the APa(Fa) exists, i.e. provided that the hull topology is Hausdorff. 4. Strict inductive limits. A topological inductive limit £[£] = Yj £<xKJ is said to be strict if Ea cz Ep for a< /? and if the topology a induced by Xp on the subspace Ea of Ep is equal to 2a. In particular, by §18,5.(2) every locally convex direct sum E[i] = © £a|jXa] is the strict inductive limit of its finite partial sums. a By § 18, 5.(2) the sum topology *% induces the topology £a on each Ea. The question of whether this is also true for the hull topology of every strict inductive limit has recently been settled by a counterexample given by Komura [2]. In the countable case, however, we have (1) Let E be the union of a strictly increasing sequence Ex\%{\ cz E2\Z^\ cz ••• of locally convex spaces £„[£„], and suppose that 2W+1 induces the topology Hn on En. Then the hull topology H is Hausdorff so that E pX] is the strict inductive limit of the spaces En pX„]; further H induces the topology Xn on each En. First we establish two lemmas. We shall only need the first of these in the proof of (1). (2) // V is an absolutely convex neighbourhood of o in a linear subspace H of the locally convex space E[i], there is an absolutely convex neighbourhood U ofo in E for which U nH=V. Using the definition of the induced topology, there exists a neighbourhood Wofo with Wc\H cz V. We can suppose that Wis absolutely convex. We form U=\~(WuV). Every zeU has the form z = ax + (ly, xe W, ye V, \a\ + \f}\ g 1. If z lies in H, x also lies in H, so that xeHn W => V. Consequently zeV and U nH =V. (3) If further, H is closed in £|JX] and x0$H, there is an absolutely convex neighbourhood U of o with Uc\H=V and x0$U. We can choose an absolutely convex neighbourhood W of o for which WnHc^V and {x0+W)nH is empty. Then if U=r(W\jV\
5. (LB)- and (LF)-spaces. Completeness 223 UnH=V, by (2), and x0 does not belong to U; for if x0 = ax + Py with xeW, yeV, |a| + |]8|^l, then x0-ax = Py, which contradicts the fact that (x0 + W)nH is empty. Proof of (1). Let Vk be an absolutely convex neighbourhood of o in Ek. By (2), there is a sequence of absolutely convex neighbourhoods of o Vk+1 cz Vk + 2 cz •••, where Vk + m is a ^Xfc+m-neighbourhood of o in Ek+m, for which Vk + mnEk=Vk. For the ^-neighbourhood of o 00 °° U = \~ Vk+m= [j Vk + m we then have UnEk=Vk. This shows that X m=1 ro=l induces the topology Xk on Ek. Every non-zero x in E lies in some Ek, and so there exists a Vk with x$Vk; x$U for a neighbourhood Uofo constructed in this way, and so the hull topology is Hausdorff. 00 (4) Let E[X] = Yj Ek[Xk~] be a strict inductive limit, and suppose fc=i that for each k Ek is a proper closed subspace of Ek + l [Xk + l~\. A subset B of £[£] is bounded if and only if it lies in some Ek[Xk~\ and is bounded there. The condition is sufficient, by (1). Conversely let us suppose that there is a bounded set B in E[X] which is not contained in any En. There then exists a sequence x{ e B and a sequence n{ with xf e En. ~ En. _ 1. Because En._ l is closed in Eni, there exists, by (3), a sequence of absolutely convex neighbourhoods Vn. ofo in En. with Vn.nEn._l = Vn._l and with 1 °° — xf$ Vn. U = l) Vn. is an absolutely convex ^-neighbourhood ofo in E i ' i=i ' 1 which contains no — xf. This contradicts the boundedness of the sequence xf. ' Thus we have proved an analogue of § 18, 5.(4), although § 18, 5.(4) is by no means a special case of (4). 5. (LB)- and (LF)-spaces. Completeness. A locally convex space E is called a (strict) (LB)- (respectively (LF)-) space if it can be represented as the (strict) topological inductive limit of a properly increasing sequence £i[Ii]c£2[I2]c" of (B)- (respectively (F)-) spaces. We shall make a deeper investigation of the properties of these spaces in the second volume; here we investigate the question of their completeness, which can be tackled in a rather more general setting. (1) Let E[X~\ be the topological inductive limit of a strictly increasing sequence £ipi]c£2[I2]c- of locally convex spaces. For each %-Cauchy filter 5 there is a coarser %-Cauchy filter 5' and a number k for which 5'n Ek is a %-Cauchy filter on Ek.
224 § 19. Locally convex hulls and kernels Proof. As F runs through a base of g and W runs through all the absolutely convex neighbourhoods of o in E\%\ the sets F+W form the base of a Cauchy filter g' on FpX], which is coarser than g. For if F is small of order W, F + W is small of order 3 W, so that the collection of all the sets F+W contains sets of arbitrarily small order. Further the set (Fx + Wx) n {F2 + W2) contains the set F3 + (W^ n W2) if F3czFxnF2. If there exists a k for which all the sets (F+W)n Ek are non-empty, then (1) is established. We suppose that this is not the case. Then there is a sequence W1^W2zd--- of absolutely convex ^-neighbourhoods of o, together with sets Fk+Wk, where Fk is small of order Wk, for which the sets (Fk+Wk)nEk are empty, for fc=l,2,.... We put W[n)= WknEn; Wkn) is an absolutely convex ^-neighbourhood of o. Consequently the sets are ^-neighbourhoods of o, for k= 1, 2,.... Let F'ke% be small of order Vk. We shall show that {F'k+Vk)nEk is also empty. Since F'knFk is non-empty, there is an x0eFk in F'k. Elements y of Fk and elements z of F'k+ Vk therefore take the form k k k y=*o+ Z <*ixi> z=*o+ Z a.*.+ Z <x'i> i = 1 i = 1 i = 1 with xt, xJeW'for i<fc, xk9 x'ke%Wk, £|a£|^l, M^L Now ttkxk + ttkxkejWk + jWk = Wk. Since x0eFfe and (Fk+Wk)nEk is empty, the element x0 + (xkxk + ukxkeFk+Wk cannot lie in Ek. On fc-l /c-l the other hand Z 0Lixi^- Z a;xi nes m ^fc-i> so tnat z a^so does not i = i i = i lie in Ffc; hence (F[+Kk)n£k is empty. 00 We now form the neighbourhood U = [~ \ W[k) of o. U <= Vk for /c=i each /c, since W{P^Wk for each /c'^/c. There is a set F0eg which is small of order U. Let y0 be an element of F0; y0 lies in some Ek. We assert that F'knF0 is empty, which contradicts the filter properties. For if yeF'k, y+Vk contains no element of Ek (for (F'k+Vk)nEk is empty) and so y — y0 does not lie in VkzD JJ\ this means that y does not belong to y0 + U, and so does not belong to F0, i.e. F'k n F0 is empty. The following completeness criterion is a simple consequence of (1): (2) Let FpX] be the topological inductive limit of a strictly increasing sequence Ft [3^] <=F2|jX2] <= ••• of locally convex spaces. E\fX\ is
6. The locally convex kernel of locally convex spaces 225 complete if and only if, for each n, every X-Cauchy filter on En has a limit in E, and so has a limit in some En + k. As a special case we have (3) Let E[%] be the strict inductive limit of the strictly increasing sequence F^^] <= F2[£2] ci ••• . E[%~\ is complete if each Ffc[£fc] is complete. In particular, every strict (LF)-space is complete. The following theorem is due to Grothendieck [13]: (4) Let A be a continuous linear mapping from an (F)-space F into 00 the (LF)-space E[X]= [j E „[_%„]. There exists a k with A(F) c Ek, n=l and A is a continuous mapping from F into Ek[_X^]. Proof. Let Hn be the set of all pairs (y,Ay)eFxEn, with AyeEn. Hn is a closed linear subspace of F x En, and so it is an (F)-space. If Pn is the continuous mapping Pn(y, A y) = y from Hn into F, Pn(Hn) is the set of all y with AyeEn. Using Baire's category theorem, it follows 00 from F = (J Pn(Hn) that some Pk(Hk) is not meagre in F. It follows n= 1 from the Banach-Schauder theorem that F = Pk(Hk), so that A(F)c=Ffc. The graph Hk of the mapping A of F into Ek is closed, so that A is continuous, by the closed graph theorem. (5) Every absolutely convex, bounded, complete subset M of an 00 (LF)-space E[X~\ = (J £„[£„] is a bounded subset of some Ffc[£fc]. n= 1 The linear subspace EM of E generated by M is a (B)-space with unit ball M (cf. § 20, 11.(2)). If we apply (4) with F = EM and A equal to the continuous embedding of EM in E, we obtain (5). Finally we remark that the question of the completeness of a topological inductive limit of complete spaces £apXa] is identical with the question of whether the quotient space (0£apXJ)/H, which by 1.(3) is topologically iso- a morphic to lim£apXJ, is complete. Since 0 £a[XJ is complete, by §18,5.(3), a any example of an incomplete lim£apXJ (with the £apXJ complete) at the same time gives an example of an incomplete quotient of a complete space. We shall give such an example in § 31, 6. 6. The locally convex kernel of locally convex spaces. There is a close parallelism between the ideas of this and the following numbers and the ideas of numbers 1. to 3. of this paragraph. Suppose that E is a vector space and that we are given a collection of vector spaces Ea and a collection of linear mappings Aa from E into Ea, such that for each non-zero x there is at least one Ea in which the 15 Kothe, Topological Vector Spaces 1
226 § 19. Locally convex hulls and kernels image Aax is non-zero. We then call E the kernel of the A(a_1)(£a), and write E=KAi~1\Eay a As an example, if the spaces Ea are subspaces of a vector space H, their intersection E = f]Ea is equal to KJ(a_1)(£a), where Ja is the a a embedding of E into £a. The product E = T\Ea can also be represented a as the kernel KP(a_1,(£a), where Pa is the projection of E onto £a. a Conversely we have (1) Every kernel E= KA{~1](E^ is isomorphic to a linear subspace EofT[Ea. a The mapping Ax = x = (Aax) of E into TT£a provides the required a embedding, because of the hypothesis that for non-zero x at least one Aax is non-zero. On the other hand, we have (2) Every linear subspace E of T\Ea can be represented as the kernel KP(a_1)(£J, where Pa is the restriction to E of the projection Pa ofT\Ea a a onto Ea. If the spaces Ea are locally convex spaces £a[£a], then, following the model of the topological product, we introduce as kernel-topology on E= KA(a~1)(Ea\_Ta]) the coarsest topology for which all the Aa a are continuous mappings from E into £a[£a]. We now determine a base of ^-neighbourhoods of o. If Ua is an absolutely convex ^-neighbourhood ofo in Ea, V(X = Ai(X~1)(Ulx) must be a ^-neighbourhood ofo. The sets Va are absolutely convex and absorbent. Their finite intersections clearly form a base of ^-neighbourhoods ofo. Since every non-zero x in E has a non-zero image Aax in some Ea, the kernel topology is always Hausdorff. E[Z]=KA(a~1)(E0i[Xa]) is a thus always a locally convex space; it is called the locally convex kernel of the ^"^(^[IJ). The algebraic isomorphism (1) can be sharpened to (3) Every locally convex kernel E[%] = K A{~ ^(i^plj) is topologic- a. ally isomorphic to a linear subspace E of the topological product TT £a[£J. a Proof. Under the mapping Ax = x = (Aax) of E onto the subspace E of TT£a, Va = A{-l\Ua) is sent to P(a_1)(^a); the sets P{~l\Ua) form a sub-base of neighbourhoods of o in E for the topology induced on E by the product topology on TT£a[£a].
6. The locally convex kernel of locally convex spaces 227 The converse is trivial: (4) Every linear subspace E[X] of a topological product E[X] = T\Ea[Xa] is topologically isomorphic to the kernel KJ*~ ^(^aPd)* ^ a ^ a where Pa is the restriction to E of the projection Pa of E onto Ea. In particular every linear subspace H of a locally convex space E[X] can be considered as KJ{~1)(E[X])i where J is the embedding of H into £[£]. If a locally convex space E[X] is represented as a kernel in two different ways, £[2] = K^-^pJ) = KB*"1^^]), we again speak of equivalent defining systems. (5) The formation of locally convex kernels is transitive. For the locally convex kernel E[X~] = K A[~ n (K £<" "(F,. [I,.])) can also be considered as the locally convex kernel K A{~ l)B{p~ ^(^VaP^J)' in either case a base of neighbourhoods ofo is formed by the finite intersections of the sets (Bp^A^'^iUpJ, where each UPa is a ^-neighbourhood ofo in FPa. (6) A linear mapping B from a locally convex space Fp'] into a locally convex kernel E[X] = KAi(X~1)(E(X[Xlx]) is continuous if and only a if AaB is a continuous mapping from F[%'~\ into £apJ, for each a. Proof. The continuity of AaB follows from the continuity of Aa and of B. Conversely if all the AaB are continuous, and if ^"^([/J is a neighbourhood of o in the sub-basis defining the topology X on £, there is a neighbourhood W of o in F with AaB(W)cz Ua, so that B{W) cz A{~ ^(l/J, and B is continuous. (7) A subset M of a locally convex kernel E[X] = KA{a~1)(E(X[Xa]) a is bounded (respectively precompact) if and only if Aa(M) is bounded (respectively precompact) in £apJ, for each a. Proof, a) Aa(M) is bounded if M is (§ 15, 6.(5)). On the other hand n if all the sets AJJM) are bounded, and if V = f] A{~ ^(l^.) is a £-neigh- bourhood ofo, then it follows from Aa.(M)^pi Ua. that Mczp-A^"1^^.), so that MapV, where p = m3.xpi. b) By § 15,6.(7), Aa(M) is precompact if M is. On the other hand if all the sets Aa(M) are precompact, then each set Aa(M) is covered by finitely many sets B$ which are small of order Ua9 and so M is covered by finitely many sets A{~l)(B{£) which are small of order ^"^(C/J. If V=A{-1)(Ua)nA<ji-1)(Ufi) then each MnA{~l){B^) is covered by finitely many sets which are small of order ^_1)(Ly, and consequently 15*
228 §19. Locally convex hulls and kernels M is covered by finitely many sets which are small of order K The corresponding assertion for a V which is intersection of n>2 sets ,4(a-1)(£/a) follows in exactly the same way. Consequently M is totally bounded, and so it is precompact. We mention one more consequence of the definition of the kernel topology: (8) The locally convex kernel of at most countably many metrizable locally convex spaces is metrizable. 7. The projective limit of vector spaces. The ideas of this number are also valid for vector spaces over arbitrary fields. Let E= K4(a_1)(£a) be the kernel of /4(a_1)(£a). Suppose that the set A of indices a is a directed set, and that for each pair a < /? there is a linear mapping AaP of Ep into Ea for which (1) Aa = AaPAp for a<jS. Again we write Aaa for the identity mapping of Ea onto itself. If af^/J, Aa(E) is therefore isomorphic to a quotient space of Ap(E). (2) // each Aa maps E onto Ea then Aap also maps the space Ep onto Ea, and we have (3) AapApy = Aay for a</?<y. For an arbitrary kernel E= K,4(a_1)(£a) such mappings AaP are not a always given in the first place. A kernel can however always be expressed in this form. For each finite subset A of the index set A form the product EA= T\ Edi and define the mapping AL from E into EA as the mapping <5eA AAx= T\ Adx. Write AX^A2 if AJCIA2, and put AA A equal to the <5eA projection of EAl onto £Al. Then AAl = AAuAlAAl and E= KAA~l)(EA). Further ALuAlAAljA^ = AAi?A3, so that (3) holds. (4) Suppose that E=KAia~1)(Ea) and that mappings Aap are given a. from Ep into Eai which satisfy (1). Then the relations (5) P* = AapPp for a<P hold for the space E=KP{a~1)(Ea), which is isomorphic to E, by 6.(1) and 6.(2). This follows directly from P0ix = A(Xx = A(xpApx = AapPpx for all xeE.
7. The projective limit of vector spaces 229 So far we have started from a vector space £, together with mappings Aa of E onto spaces Ea and mappings AaP of Ep into Ea. To what extent is E determined by the spaces Ea and the mappings Aap alone? (6) Suppose that Ea, cceA, is a directed system of vector spaces, that a linear mapping Aap from Ep into Ea is given for each cc < /?, and that (7) AapApy = Aay for cc<p<y. We denote by E the linear subspace of E = T\Ea= KP(a_1)(£a) consisting of all x = (xa) with xa = AaftXp for a</J. a a // E is a linear subspace of £, and if Pa is the restriction of Pa to E, then E= KP(a_1)(£a) and (5) is satisfied. a This follows from 6.(2) and the hypothesis that x(X = AapXp for the components of the elements of E. E is the largest possible subspace E of E. The uniquely determined space E= KP(a_1)(£a) defined in this way is called the projective a limit of the space Ea under the mappings A^, and it is also denoted by lim Aap(Ep). It can happen that lim Aap(Ep) reduces to the single element o. In contrast to 2.(10), where it follows from the fact that the mappings Aap are one-one that the mappings X(a0) have the same property, if E is a projective limit for which the mappings Aap map Ep onto Ea it need not follow that the Pa map E onto Ea. This is so, however, in the following special case: (8) Suppose that the directed set A is countable. If the Aap, cc < /?, ae A, each map Ep onto Ea, then in the projective limit E= KP^^EJ the Pa map E onto Ea. a In this case we can therefore be sure that £=#o if the Ea are different from o. Proof. We must show that given x^eEp there is an element in E whose component in Ep is x(p0). We arrange the indices a as a sequence af, i=l,2,..., with <x1=fi. If <Xj<P, we put x^^A^pX^K Let aix be the first term in the sequence of indices at for which ai<^(x1, and let akl be the first at with af^afl and ^xx^ By hypothesis there is an x^} with xf^A^x^. For all a,<afcl we again put x^ = Aaj^ki x[°J: Because of (7), this is consistent with the terms which have already been fixed. We now repeat the procedure, letting aI2 be the first af with 0Li^takl and cck2 the first a, with af^aI2 and af>akl, and so on. In this way we obtain an element x(0)eE with component x^ in Ep.
230 §19. Locally convex hulls and kernels (9) If E — K^~ ^(EJ, and if mappings Aa0 are given from E0 into Ea a which satisfy (1) and (3), then E is always isomorphic to a linear subspace of lim AaP(Ep). We have only to go over from E to the space E= K P{a~ ^(EJ, which a is isomorphic to it by 6.(1) and 6.(2), to obtain a linear subspace of Urn \piEpl using (4) and (6). 8. The topological projective limit of locally convex spaces. If the spaces Ea are locally convex spaces £a|jXa], and if the mappings Aap from Ep into Ea are all continuous, then on E = limAap(Ep[_Xp]) we define as topology X the topology induced on E by the topology of TIEplZp], i.e. the kernel topology of KP<_1)(Ea[3;a]). E[X] is called P a the topological projective limit of the Aap(Ep[Xp]). (1) // E[X] = KAi~1)(Ea[Xa]) is a locally convex kernel, and if a continuous mappings Aap are given from Ep into Ea which satisfy 7.(1) and 7.(3), then E[X] is topologically isomorphic to a linear subspace E of the topological projective limit lim Aap(Ep[Xp]. A base of X-neighbourhoods of o is given by all sets of the form Va = A(a~l)(UlxX where Ua is a Xa-neighbourhood of o in Ea[Xa~]. Proof. The first part of the assertion follows from 7.(9) and the fact that the topologies on E and lim Aap(Ep) are the kernel topologies induced by the product topology on TT£a[^J. a A ^-neighbourhood of o for the kernel topology can be taken in the n form V= f] A{~l\Ua^ where Ua. is a ^.-neighbourhood of o in Ear i=i Since the indices form a directed set, there exists a /? with at<jS for i=l,...,M, and there exists a U^aEp with Aa.p(Up)czUar It follows from this, and from 7.(1) and 7.(3), that A}T "(l/,) ci A[~ "A^AJT l)(Up)) = A[~ "A^A,(A!f l\Up)) = ^(«71)^(^)c^"1)(i/J, i.e. ^~1)(^)c: K and so tne second part of the proposition is proved. It follows from the remark after 7.(2) and from (1) that every locally convex kernel is topologically isomorphic to a linear subspace of a topological projective limit. In particular the topological product TT£a[£a] can be represented a as the topological projective limit limPAl A2(£A2pAJ) of the finite partial products EL[X^]= Tl Ed[Xd~] under the projections PAl)A2 of £A2 onto £Al. SeA
9. The representation of a locally convex space as a projective limit 231 (2) Let E be the topological projective limit E[X] = limAap(Ep[%p]) of locally convex spaces £a[£a], and let V be a cofinal subset of the directed index set A. As y and 6 run through all pairs y<5ofV, E+ = HmAyd(Ed[Xd~]) is topologically isomorphic to E[X]. Proof. E consists of all x = (xj, aeA, with xa = AapXp for a</}. If each xeE we make correspond the element x+ = (xy), yef of E + , we obtain a homomorphism of E into E+. Conversely suppose that x+ =(xy) belongs to E+. For each aeA there is a yeT with a<y; we set xa = Aayxr Because of 7.(3), xa is independent of the choice of y, and for the same reason we have xa = Aapxp for each pair a</?. Consequently each x+eE+ corresponds in a one-one way to an xeE, so that E and E+ are isomorphic. This is a topological isomorphism: by (1), we can restrict our attention to neighbourhoods U'a of o, which consist of all x = (xa)eE for which xae£/a, where Ua is a neighbourhood of o in Ea. If a<(5el~, U'a also consists of all x with xdeA\^ ^(UJ. To this there corresponds the neighbourhood of all x+=(xy)e£ + with x5eA^l)(lJ^ and so the isomorphism between E and E+ is a topological one. 9. The representation of a locally convex space as a projective limit. We return to the topological isomorphism x->x = (xa) between a locally convex space E[X] and a subspace £ of a topological product of (B)-spaces which was established in § 18, 3.(7). (1) Every locally convex space E[%~\ is topologically isomorphic to a dense linear subspace of a topological projective limit of (B)-spaces. Every complete locally convex space is isomorphic to a topological projective limit of (B)-spaces. Proof. We take a system {pa(x)} of semi-norms on E corresponding to a base of ^-neighbourhoods of o. The indices a form a directed set when we define a^jS if pa(x)^pp(x) for all xeE. If we again denote the coset of xeE in Ea = E/Na (respectively Ep = E/Np) by xa (respectively kp\ then a continuous linear mapping Aaf} is defined from the normed space £/?[£/?] onto the normed space £a[£j by setting xa = Aa£xp. This mapping can be extended to a continuous linear mapping Aap from the completion JE^p^] into £a[£a]. The relation AapApy = Aay is satisfied for a<jS<y, and consequently we have AapApy = Aay. As a result, we can form the topological projective limit E[Z] = \imAap(Ep[Xp])=KP{~1)(Ea[±J), where Pa is the restriction to E of the projection Pa of T7£a[Ia] onto Ea. Under the mapping a x-+x = (xa), E[X~\ is topologically isomorphic to the subspace E of all such elements x in £, since x(X = A(XpXp if a</J.
232 § 19. Locally convex hulls and kernels We now show that E is dense in E. By the second part of 8.(1) this is so if Pa(E) is dense in Pa(£) for each a. But Pa(E) = Ea, and Pa(E) c Ea. As a special case, we have (2) Every (F)-space is topologically isomorphic to a projective limit of a sequence of (B)-spaces. 10. A criterion for completeness. In contrast to the behaviour of inductive limits (cf. 5.), the question of the completeness of a projective limit is easy to answer. Let E\Z~\ be a locally convex kernel KA{~ l)(Ea\X^. a (1) A filter 3 on E\%\ is a Cauchy filter if and only if all the filters i4a(g) are Cauchy filters in the spaces £a[£a]. Suppose that 3 is a Cauchy filter. The continuous images Aa(g) are Cauchy filters in the spaces £a[£j. Conversely, suppose that each n v4a(5) is a Cauchy filter in £a[£a]. Suppose that V= f] A{-l\Ux) is a i= 1 ^-neighbourhood of o in £[£], where Ua. is a ^.-neighbourhood of o in £aipXaJ. By hypothesis there is an Ffeg whose image Aai(Ft) is small n of order Uar Then Ft is small of order A(a_1)(C/a.) and f]Ft is small of order V, so that g is a Cauchy filter in E[%~\. i= 1 For arbitrary locally convex kernels the completeness of E[Z~] certainly does not follow from the completeness of the spaces EapXJ. This follows from 6.(4), taking the £a[£J to be complete and E[X~\ to be a non-closed subspace of £[3f). For projective limits, however, we have (2) A topological projective limit E[X] =lim Aap(Ep[%p]) is complete (respectively quasi-complete) (respectively sequentially complete) if this is so for each space £a[£a]. Proof. In the terminology of 7.(6) we have £[£] = £[£] = KP(a_1)(£a[Xa]). Suppose that the spaces £apXa] are complete, and let g be a Cauchy filter on E[Z]. Then each of the Cauchy filters Pa(g) has a limit xa in £a[£j. If a< jS, it follows from 7.(5) and the continuity of AaP that Xx = \imPa(%) = \imA^Pp(%) = A^\imPp(%) = AaPxp. The element x = (xa) therefore satisfies the compatibility conditions of 7. (6) and so it belongs to E = E. It follows directly from the fact that Pa(g)-+xa for each a that %-+x in £, so that £[£] is complete. If 3 is a Cauchy filter on a bounded set M czE[X], the Pa(%) are Cauchy filters on the bounded sets Pa(M) in £apXj, and we finish the argument as before.
1. The existence of continuous linear functional 233 The assertion about sequential completeness is obtained in a similar way. (3) A topological projective limit E[%~\ =lim Aap(Ep[Xp]) is a closed linear subspace of the topological product TTFa[£a]. a For a closure point of F[£] in TTFa[£a] is the limit of a Cauchy filter in FpX], and so, as in the proof of (2), it satisfies the compatibility conditions of 7.(6); consequently it belongs to FpX]. References to additional results on inductive and projective limits: N. Bour- baki [3], vol. 3., J. Braconnier [1], S. Lefschetz [1], D. A. Raikov [1], J. Se- BASTIAO E SlLVA [4], O. TAKENOUCHI [1], A. WEIL [2]. § 20. Duality 1. The existence of continuous linear functional. In § 17,6. we showed, using the Hahn-Banach theorem, that there are sufficiently many continuous linear functionals on any normed space F, and derived results about the dual space E'. It is also possible to do this in the more general locally convex case. Once again, the extension theorem holds: (1) Every continuous linear functional l(z) defined on a linear subspace F of a locally convex space E[%~\ can be extended to a continuous linear functional u defined on the whole of E. In particular if \l(z)\^p(z) on F, where p(x) is a continuous semi- norm on F[£], there is an extension u with \ux\^p(x) on the whole of F. Proof. If l(z) is continuous on F, then by § 18,1. there exists a continuous semi-norm p(x) on E with \l(z)\^p(z) on F. The existence of a continuous extension u of /, with |wx|^p(x), follows from § 17,3.(5). (2) // xl5...,x„ are linearly independent elements of E[X] and if a!,..., a„ are real (respectively complex) numbers, there exists a ueE' with uxi = oihi=\,...,n. In particular, for each non-zero x0 in E there exists a u0eE' with Proof. The ^-dimensional subspace F of E spanned by x^,..., xn is topologically isomorphic to K", by § 15, 5.(1). The equations l(xt) = ah i = l,..., w, therefore define a continuous linear functional on F, and this has a continuous extension u to the whole of F, by (1). A third theorem on the existence of continuous linear functionals which is frequently useful is given by (3) // F is a closed linear subspace of E[X] and if x0$F, then there is a u0eE' with u0x0=l and uoy = 0 for all yeF.
234 §20. Duality Proof. E/F is again locally convex under the induced topology X, by §18,3.(2). By (2) there is a continuous linear functional u0 on E/F with u0x0 = 1, where x0 is the coset Kx0 of x0 in E/F. A linear functional u0 is defined on E by setting u0x = u0x, where x is an arbitrary element of E; clearly w0x0 = l and u0y = u0o = 0 for all yeF. w0, being the product of the continuous mappings K and w0, is a continuous linear functional on E. 2. Dual pairs and weak topologies. We now use the concept of dual pair, defined in § 10, 3., once again. (1) // E\%\ is a locally convex space, E and its dual space E' form a dual pair <£',£) over the real (respectively complex) field. Proof. As in §10,3. we define B(u,x) = ux, i.e. the value of the linear functional u at x, as the bilinear functional on E' x E. Condition (D2") is trivially satisfied, and condition (D2') follows from 1.(2). Thus we have established a result for locally convex spaces which we obtained for linearly topologized spaces in § 10, 4. At the same time, the results which we proved in Chapter II for dual pairs over arbitrary fields are also valid for dual pairs consisting of a locally convex space and its dual space. As in Chapter II, it will become apparent that many of the most important concepts of the duality theory of locally convex spaces do not depend upon the original topology X of E, but only on the dual pair <£',£>. As in §10,3. we shall now start from a dual pair <E2>£i> over K, where now K can only be the field of real or complex numbers. We shall denote the bilinear form by ux, or occasionally by <w,x>. In §10,3. we introduced a linear topology on E1 which we called the linear weak topology Xls. As we saw in § 10,4.(4), this is the coarsest linear topology on E1 for which all the ueE2 are continuous linear functionals on Ex. For this, the topology on K is the discrete topology. For our study of topological vector spaces in Chapter III, however, the underlying topology on K is the topology defined in terms of the modulus. For this reason, we now define the weak topology XS(E2) on Er to be the coarsest topology on E1 for which each element ueE2 defines a continuous linear functional <u,x>=wx on El9 when K is given the topology defined in terms of the modulus. The sets UUl Un.E form a base of neighbourhoods of o for the weak topology XS(E2), where UUl,...,u„;£ is the set of all xeEx for which (2) sup |M,-x|<e; i= 1,..., n the u( are n arbitrary elements of E2, and n= 1,2,.... For if ueE2, the set UU;E of all xeEx for which |wx|<a must be a neighbourhood ofo.
2. Dual pairs and weak topologies 235 On the other hand every intersection of neighbourhoods of o of this form contains a neighbourhood of o of the form (2). The topology 2s(£i) is defined on E2 in a similar way. As in § 10, the following result holds for these topologies: (3) // (E^E^ is a dual pair, El[Zs(E2)~\ and E2[Zs(El)] are locally convex, and each is the dual of the other. The proof proceeds in a completely analogous way: the neighbourhoods (2) are absolutely convex and absorbent and define a filter, and as a result of 1.(2), the weak topology is Hausdorff; thus El[Zs(E2)] and E2[2S(E1)] are locally convex, by § 18,1. Each ueE2 defines a weakly continuous linear functional on El9 since |wx|<£ for each xeUu;E. Conversely if u is a weakly continuous linear functional on E± there is a weak neighbourhood l/ = t/Ml,...,Mn;fi of o with |mx|^ sup \utx\. i=l,...,n In particular ux = 0 if utx = 0 for i=\,..., n. But then by §9, 2. (7 a) u is a linear combination of the uh so that ueE2. If E[Z~] is a locally convex space, then using the dual pair <£',£> we can introduce the weak topology ZS(E) on E, and we can likewise introduce the weak topology XS(E) on the dual space E; this makes E a locally convex space. We call E[ZS(E)~\ the weak dual of E[2]. The relation between the original topology 2 and the weak topology is given by (4) The original topology 2 on a locally convex space E[%~] is always finer than the weak topology 2S(£')- For every element of E is 2-continuous, while ZS(E) is the coarsest topology for which this is so. If E is an infinite-dimensional normed space, the norm topology is certainly different from the weak topology, since none of the semi-norms sup \u(x\ is a norm. We make the following further remark about the relation between locally convex spaces and their corresponding dual pairs (cf.*§ 10,3.): (5) // the locally convex spaces £i[2i] and E2[Z2~\ are isomorphic, then the dual pairs (E^E^ and (E'2,E2} are isomorphic; the converse need not be true. The fact that the converse is not true follows from the fact that the original topology cannot be retrieved from the dual pair. If two dual pairs are isomorphic, however, it is at least possible to deduce that the spaces are topologically isomorphic when they are given their weak topologies. For these topologies depend on the dual pair alone.
236 §20. Duality 3. The duality of closed subspaces. Let <£2,£i> be a dual pair. We recall the concept of the space M1 orthogonal to a set Mc£1? which was introduced in §9,2. M1 consists of all ueE2 for which uy = 0 for all yeM. Just as for linear topologies, the following theorem also holds for locally convex topologies: (1) If F is a Z-closed linear subspace of the locally convex space £[2], F is orthogonally closed with respect to E; conversely a linear subspace F of E which is orthogonally closed with respect to E' is 2S(£')- closed in E. The proof proceeds in a way analogous to the linear topology case (§10,4.(6)): it follows from 1.(3) that F is orthogonally closed if it is ^-closed. On the other hand if F1A- = F, if x0 is a weak closure point of F and if UqeF1, then there is at least one y0 eF in each weak neighbourhood Uu;£(x0). Consequently \u0x0\ = \u0(x0 — y0)\<s, so that, as e is arbitrarily small, woxo = 0, and x0eF11 = F. We have thus shown that although the original topology 2 of a locally convex space E[Z~\ is certainly not determined by the dual pair <£',£>, nevertheless the 2-closed linear subspaces are determined by the dual pair alone. If we call a locally convex (respectively linear) topology on Et compatible with the dual pair <£2,£1> when the dual space of Et is E2, then we can formulate this result as follows: (2) // <£2, Ety is a dual pair, a linear subspace F of Et is closed with respect to a compatible locally convex or linear topology if and only if it is orthogonally closed. In this sense we can thus speak simply of "closed" linear subspaces. The question of characterising all the compatible linear topologies was answered in §10,11.(4), and a characterization of all the compatible locally convex topologies will be given in §21,4.(3). So far at any rate we know one compatible topology, namely the weak topology. If we define /\Fa to be the intersection f]Fa of the closed linear a a subspaces Fa of E{ and define \/ Fa to be the smallest closed linear sub- a space of Et containing all the Fa, then the closed linear subspaces of Et form a complete lattice K(£i), and the duality properties of closed linear subspaces follow from § 10, 3.(2): (3) Let {E2,E{} be a dual pair. The complete lattices K(Et) and V(E2) of closed linear subspaces of Ei and E2 respectively are dually isomorphic; a dual isomorphism is obtained when we make each such subspace correspond to its orthogonal space.
4. Duality of mappings 237 In particular this duality holds for the ^-closed linear subspaces of a locally convex space F[2], and the 2s(F)-closed linear subspace of the dual space E'. This duality of closed linear subspaces will prove in 8. to be a special case of a more general duality. As in § 14, 7. a subset M c F[2] is said to be total in £[2], or a fundamental subset of F, if the closed linear span of M coincides with E. We clearly have (4) A set M is total in E\%~\ if and only if it is total with respect to any locally convex or linear topology compatible with the dual pair <F', F>, i.e. if and only if M1=o. Thus the concept of total set depends only on the dual pair, as well. 4. Duality of mappings. Suppose that two dual pairs <E2,£1> and <F2,F1> are given. Then by § 10, 3., E2 and F2 are linear subspaces of the algebraic dual spaces Ff and F? respectively. To every linear mapping A from E{ into F{ there corresponds the adjoint mapping A' from Ff into Ff, defined by v(Ax) = (A'v)x for all xeF1? veF^. In what follows, A' will always mean the restriction of A' to F2 c F%. (1) Suppose that (Fjj^i) and ^2^) are two dwa/ pairs ewer K. y4 /mear mapping A from Ft into Ft is weakly continuous if and only if the adjoint mapping maps F2 into E2. Corollary. A is weakly continuous if and only if A' is weakly continuous. Proof, a) Suppose that A'(F2)czE2. If U=UVly^>Vn.E, vteF2, is a weak neighbourhood of o in Fl5 then, since vi(Ax) = (A'vi)x and since A'v{eF2, the image Ax of an xeUA>Vx,...jA>Vnie is contained in U. Thus A is weakly continuous at o, and consequently it is weakly continuous on Et. b) Suppose that A is weakly continuous. Then the linear functional on Ft defined by l(x) = v0(Ax), v0eF2, is weakly continuous, since v0 is weakly continuous, by 2.(3). l(x) is therefore defined by some u0eE2; on the other hand, v0(Ax) = (A'v0)x, so that A'v0 = u0eE2, and A' maps F2 into F2. c) If A is weakly continuous, then A' maps F2 into F2, by b). The mapping (A')' adjoint to A' is equal to A, and so maps Ft into Ft. By a), A' is therefore weakly continuous. Since A" = A, the corollary follows by interchanging A and A'. We observe that this method of proof has already been used in § 10, 5.(1); accordingly, (1) can also be derived for the linear weak topologies in exactly the same way. Consequently
238 §20. Duality (2) The Zs-continuous and Xls-continuous linear mappings from a locally convex space E[Z~\ into a locally convex space F[%'] are the same. If £i[2i] and £2[£2] are two locally convex spaces, we denote the vector space of all continuous linear mappings from Et into E2 by 2(E1 [21],E2[22])- The following duality theorem now follows from (1): (3) Let <£2»^i) and (,Fi^F\) be two dual pairs over K. // to each weakly continuous linear mapping A from El into Fx we make correspond its adjoint mapping A' from F2 into E2, we obtain a vector space isomorphism between the spaces 2(El[Zs(E2)~\, Fl[Xs(F2)]) and 2{F2[Xs(Fl)~\, E2[2s(El)-]). In the case where the two dual pairs coincide, we obtain (4) The correspondence which sends each weakly continuous endomor- phism A of E{ into the weakly continuous adjoint endomorphism A' of E2 maps the algebra Q(El [2S(£2)]) °f weakly continuous endomorphisms of E{ anti'isomorphically onto the algebra fi(£2ps(E1)]) of weakly continuous endomorphisms of E2. In 2. we found that every continuous linear functional on a locally convex space is weakly continuous, and conversely. In general, for linear mappings we can only say (5) Every continuous linear mapping A from El[Zl~\ into E2[Z2~\ is also weakly continuous, i.e. ^(E^X^, E2[X2~\) is a linear subspace of fi^p^Ei)], £2[£s(E2)]). Proof. The linear functional defined by l(x) = v0(Ax), v0eE'2, is X!-continuous, and so it is defined by some u0eE\. A' therefore maps E2 into E\, so that A is weakly continuous, by (1). E be an infinite dimensional normed space. The norm topology X is finer than, and different from, the weak topology XS(E'). The identity mapping of E onto itself is certainly an element of £(£[XJ, £[XJ), but does not belong to £(£[XJ, £[X]). From (5) and the corollary of (1) we obtain (6) // A is a continuous linear mapping of E^IJ into £2[£2], A' is a weakly continuous linear mapping of E2 into E\. 5. Duality of complementary spaces. In §15,9.(10) we saw that in arbitrary topological vector spaces not every finite-dimensional linear subspace need have a topological complement. Locally convex spaces have this property, however, as do linearly topologized spaces (cf. § 10, 7.(8)); this we shall now show. Every weak topological complementary decomposition has a dual decomposition:
5. Duality of complementary spaces 239 (1) Let <£2»^i) be a dual pair. To every % ^complementary decomposition (2) E1=H1®H2 of E{ into two closed linear subspaces there corresponds the ^-complementary decomposition (3) E2 = H\®Hl If Px and P2 are the two projections corresponding to the decomposition (2), with Pl(El) = H\,P2(El) = H2, then their adjoints P/ and P2 are the projections corresponding to the decomposition (3), with P[{E2) = H2,P2(E2) =H\. Proof. A weakly continuous projection of Et is a weakly continuous endomorphism P of El9 with P2 = P. By 4.(4) the adjoint mapping P' satisfies the equation (Pf)2 = Pf, so that P' is a weakly continuous projection of E2. The projections corresponding to (2) satisfy the equations J = Px + P2, Pt P2 = P2P1= 0. Again by 4. (4), the adjoint mappings satisfy the equations /=p;+f2, p'2p[=p[p2=o. It follows from §15,8.(1) that P[ and P2 determine a ^-complementary decomposition of E2, and E2 = P[(E2)@ P2(E2). We still have to show that P[{E2) = H2l and P2(E2) = H^ P[{E2) consists of all ueE2 with P[u = u. This is equivalent to (Plu)x = u(Plx) = ux = u(Plx + P2x) = u(Plx) + u(P2x) for all xe£l5 i.e. ueH2. We show that P'2(E2) = H\ in the same way. Since every ^-continuous projection of a locally convex space E[Z~] is also weakly continuous, by 4.(5), we have (4) Every X-continuous complementary decomposition E = Hl@H2 of a locally convex space E[X] is also a <Xs-continuous decomposition, and it determines a dual <Xs(Ef)-continuous complementary decomposition E' = H\®Hi In general, however, it is too much to exspect a weakly continuous complementary decomposition of E\1L\ to be 2-continuous as well, for not every weakly continuous projection need be ^-continuous. (5) Let E\X\ be locally convex. Every finite-dimensional linear subspace H of E has a %-complement. Proof. H is a closed subspace of E (cf. §15,5.(2)) and, applying § 9, 2. (7 a) to E' instead of to E, its orthogonal complement H1 has co- dimension n in £', where n is the dimension of H. By § 15, 8.(2), H1 has a 2s-complement G of dimension n. Since H11 = H, to the complementary
240 §20. Duality decomposition E' = H1@G there corresponds the ^-complementary decomposition E = H © G1, by (1). But this decomposition is also ^-complementary, by § 15, 8.(2), since H is a finite-dimensional algebraic complement of G1. In § 15,12.(6) we showed that two algebraically complementary closed linear subspaces of a complete metrizable space, and in particular of an (F)-space, are always topologically complementary. This is by no means true for arbitrary locally convex spaces. In § 10.(8) we gave a simple example of a dual pair in which an algebraically complementary decomposition into X/s-closed subspaces need not be X/s-continuous. The proof also holds for the weak topologies, because of 3.(2), and because a Xs-continuous projection is also a X/s-continuous projection. By the same token, the example of a closed linear subspace of (paj®a><p with no topological complement given in §13,6. also provides an example of a locally convex space with a weakly closed linear subspace with no Xs-complement. In § 31 we go into the question of the existence of a topological complement in more detail. 6. The convex cover of a compact set. We now link up with the ideas of§ 15,6. From the fact that every locally convex space has a base of neighbourhoods of o consisting of absolutely convex sets, it follows immediately that (1) If M is a bounded subset of a locally convex space, the closed convex cover C (M) and the closed absolutely convex cover |~~(M) are again bounded. We now investigate the corresponding question for compact and precompact subsets of a locally convex space. We begin with a counterexample. The space cp of all finite vectors is a subspace of the Hilbert space/2, and cp is a normed space under the norm induced by the norm of I2. The set M of vectors x{n) = — e„ is a sequence n tending to zero in /2; together with o it forms a compact subset of cp. oc i k If a„>0, ]£ a„=l, the sequence n(/c) = -—Xa"*("} is a Cauchy n= 1 V^ 1 1 sequence in C (M), whose limit is not a finite vector and so does not belong to cp. Thus C (M) is not compact. In contrast to this, for precompact sets we have (2) // M is a precompact subset of the locally convex space E[X], C (M) and \~(M) are again precompact. It is sufficient to show that T(M) is totally bounded, for then ]~{M) and C (M) c F(M) are precompact.
6. The convex cover of a compact set 241 Let V be a closed absolutely convex neighbourhood of o. There are m finitely many points x(eM, j=1,...,w, such that M cz (J (xt + jV). i= 1 Let 5 be chosen so that SM, and consequently \~(SM\ is contained in \ V, and let A be the subset of the m-dimensional space Km consisting m of all a = (a1,...,aWI) with £ |af|^l. There is a finite set of points jg<J) = (/?</>,..., ptf) in A with the property that for each ae A there is a /?0) m with X la,.-^!^. It is sufficient to show that P(M)<= (J( £ $-/)xf+ K J. An element N j\i=l / of r~(M) has the form 3; = ^7^.3;^ with £|yfc|^l, ykeM. For each yk 1 iv iv there is an xik with yk = xik + zk, zke$V, so that )> = X>fcxik + Xykzfc m 1 1 1 m m But then for a suitable /?0), 3; = Y,fiJ)xi + 'L(<*i-PP)**2 m 1 1 = £#j)X; + z' + z with z'er(SM)cz^V, so that ye^^^+K and sor(Af)<=U(Z/^^+^) From (2) there follows (3) // M is a compact subset of the locally convex space E[%~\9 C (M) (respectively |~~(Mj) is compact if and only if C (M) (respectively |~~ (M)) is complete. This is always the case if E[X] is quasi-complete. Proof. If C (M) (respectively P(M)) is complete, then every closure point of the precompact set C (M) (respectively |~"(M)) lies in this set, and so it is compact. The second part of the assertion follows from the fact that every precompact set is bounded, by § 15, 6.(6). Later on we shall prove Krein's theorem, which extends the second part of (3) in a profound way (cf. § 24, 5. and 6.). (4) The circled cover of a compact subset K of a topological vector space is again compact. Proof. The circled cover of K consists of all ax, xeK, |a|^l. If A is the set of all aeK with |a|^l, then the topological product AxK is compact, by Tychonoff's theorem. The mapping which sends each 16 Kothe, Topological Vector Spaces 1
242 §20. Duality (oc,x)eAxK to the element axe£p] is continuous, and maps A x K onto the circled cover of K; this is therefore compact. (5) // Kl,..., Kn are convex (respectively absolutely convex) com- n pact subsets of the locally convex space E[%~], the convex cover C Kt (respectively absolutely convex cover |~~ KJ is again compact. n Proof. Let A be the set of all a = (a1,...,a„)e K" with £af = l, a^O. ; = i A is a compact subset of K". The topological product AxK{x -• xKn is again compact, by Tychonoff's theorem. If we map each n (<x,xi,...,xr^eAxKl X"-xKn to the corresponding element £afX;eE[2T|, n i=l we obtain a continuous mapping whose image C Kt is compact. i= 1 For absolutely convex sets the proof is similar; instead of A we n consider the set B of all j8 = (]81,..., j8w) with ^IP^l, fteK. i= 1 In the case of a real locally convex space we have the further result: (6) The absolutely convex cover \~(K) of a convex compact set K is compact. Proof. All elements of the form px — oy with x and y in K, p^o, o-^o and p + o-^1 lie in \~(K). On the other hand, all the elements of \~K are of this form: to show this, it is sufficient to show that if two elements zl=plxl — <rly1 and z2 = p2x2—o2y2 have this form, then so does a1z1+a2z2, where |a1| + |a2|^l. Since — z = ay — px is again of the required form, we may suppose that oc1>0 and a2>0. But then ((XiPi &2P2 a1z1+a2z2 = (a1p1+a2p2) *i + *2 (oc1al oc2a2 -(al(J1+a2G2)\-j-x1 +-J~X: where alpl +oi2p2 = c, alal -\-oc2a2 = d. Since O^c + d^l, and since the elements in the brackets lie in K, the assertion now follows. If now C is the set of all (p,a) with p^O, ff^O, p + a^l, C is compact. The mapping from CxKxK into E which sends the element (p,a,x,y) to the element px — oy is continuous, and so \~~ K is compact. For complex locally convex spaces, (6) is not true in general; however in this case we have (7) The absolutely convex cover \~ K of a compact convex set M is relatively compact, so that \~~ K is compact.
7. The separation theorem for compact convex sets 243 Proof. By (6), we can suppose that K is real absolutely convex. The set K{=K + iK is convex, by §16,1.(3), and is compact, by §15,6.(8). K{ contains every set of the form ei(pK, and so it contains the complex circled cover of K. Since Kx is convex, Kt also contains the convex cover of the circled cover of K, i. e. r~ (K), so that \~~ (K) is relatively compact. Consequently (5) can be sharpened to give (8) // £[£] is real locally convex and if Xj,..., Kn are convex com- n pact subsets of E[%~], then |~~Kt is compact. i= 1 // E[_%~] is complex locally convex and if K{,..., Kn are convex com- n ~n pact subsets of £[£], then |~ K{ is relatively compact, so that |~~ Kt is compact. I_1 I_1 The real case follows from (5), when we replace the Kt by their absolutely convex covers, which are compact, by (6). In the complex case we apply (5) to the sets \~(Kt), which are compact, by (7), to obtain that |~~ (\~K() is compact, so that f~ K( is relatively compact. i=l i=l Example. In co, the topological product of countably many one-dimensional complex spaces, the set K of all x = (xn) with xn real and |xj^ 1 is real absolutely convex, and is compact. \~~(K) consists of all y = (yn) with yn complex and \yn\^ 1. We shall show that [~~ K is a proper subset of [~ K, and is therefore not compact. The elements of \~ K have the form fc=l A coordinate fc=l fc=l with £l/?fcl = l is of modulus 1 only if all the /?fc vanish except for one /?fc=l. An element i)0 = (ei<Pn) of r^ with infinitely many different cpn can therefore never be a 3. 7. The separation theorem for compact convex sets. The existence theorems obtained in 1. were sufficient to derive the duality properties that we have obtained so far. We need a further existence theorem, however, for the duality theory of closed convex sets which we shall develop in the next number; this we shall first formulate geometrically as a separation theorem for compact convex sets: (1) Let A be a closed convex subset of a locally convex space £[£], and let K be a compact convex set which is disjoint from A. Then there exists a closed real hyperplane which separates A and K strictly.
244 §20. Duality Proof. By §15,6.(9) there exists an absolutely convex open neighbourhood U of o for which A + U and K + U are disjoint. By § 17,1.(4) these two open convex sets are separated strictly by a closed real hyper- plane, and consequently so are A and K. The analytic form follows by characterising the real hyperplane by an equation $l(ux) = y (cf. § 16, 3.): (2) Under the same hypotheses as for (1), there exists a continuous linear functional u0eE' and a real number y for which (3) sup9{(w0j;)^y< inf9l(w0z). ysA zeK The strict inequality in (3) follows from the fact that the infimum is attained in K. If instead of §15,6.(9) we apply theorem §15,6.(11), then in place of (1) we obtain (1') Let A be a complete convex subset of the locally convex space E[X~\, and let K be a closed precompact convex subset which is disjoint from A. Then there exists a real closed hyperplane which separates A and K strictly. Every closed convex 2-body C in a topological vector space is the intersection of the closed half-spaces, containing C, which correspond to the closed supporting hyperplanes of C (§ 17, 5.). The corresponding result for compact convex subsets of a locally convex space now follows from (1): (4) In a locally convex space, every compact convex set K is the intersection of the closed real half-spaces containing K which correspond to the supporting hyperplanes of K. Proof. If x0$K, then by (1) there exists a hyperplane ^R(u0x) = y which separates x0 and K strictly. Since K is compact, the infimum p in (3) is attained for some element z0eK. Then z0 lies in the hyperplane 9l(w0 x) = inf 9l(w0 z) = p, and K is contained in the half-space ^R(u0x) ^ p; zeK thus the hyperplane is a supporting hyperplane at z0 and it separates K from x0. The assertion follows from this. For arbitrary convex sets we have (5) Every closed convex subset C of a locally convex space E[X~\ is the intersection of the closed real half-spaces containing it. We have only to apply (1) to C and a point x0 which does not lie in C, since x0 is certainly a compact convex set. As a result we have shown that forming the closed convex cover of a set M depends only upon the dual space, and does not depend upon the original topology:
8. Polarity 245 (6) The closed convex cover C (M) of a set M in a locally convex space E[%~\ can be obtained using any locally convex topology compatible with the dual pair <£',£>, e.g. the weak topology. Any compatible linear topology can be used as well. As a consequence of (6) we have (7) Suppose that £[2] is a metrizable locally convex space. If a sequence xneE has x0 as a weak closure point in E, then there is a sequence of finite linear combinations of the xn which converges to x0 with respect to %. For x0 belongs to the weakly closed convex cover of the set of x„, and the convex cover of the xn is dense in this set with respect to the topology % by (6). In §17,5. we saw that (5) need not hold in arbitrary topological vector spaces. We mention one further result about compact sets: (8) Suppose that M is a compact subset of a locally convex space £[£], and that H is a real closed hyper plane in E. Then M has a supporting hyperplane parallel to H. If H is given by $R(wx) = 0, and if the maximum a of $R(wx) on M is attained at z0, then $R(wx) = a is a support hyperplane of M, since it contains z0 and since M lies in the half-space $R(wx)^a. 8. Polarity. Suppose that we are given a dual pair (i^,^). If M is a subset of El9 we call the set of all ueE2 for which $R(wx)fg 1 for all xeM the polar M° of M in E2. The polar in Ex of a subset of E2 is defined similarly. If M is circled, M° is also equal to the set of all u for which \ux\ ^ 1 for all xeM: for if |a|:g 1, ax is also in M, and it follows from $R(awx)5^1, for all |a|^l, that |wx|^l; and clearly if |wx|^l then $R(wx)^l. If we call the set of all ueE2 for which |wx|^l for all xeM the absolute polar ofM, then the absolute polar of M is the polar of the circled cover of M. If M is absolutely convex, M° is the absolute polar of M. The absolute polar is always absolutely convex. The next result follows easily from the definition: (1) The following rules hold for forming polars: 1 a) (aM)° = — M° for each non-zero ae K. a b)//Mc JV, then N° => M° and M°°c Noc. c) MczM00. d) M° = M000.
246 §20. Duality We shall only prove d): if we apply c) to M° we obtain M° <= M°°°, and if we apply b) to c) we obtain M° => M°°°; the two together give d). M°° is also called the bipolar of M. Similarly the absolute polar of the absolute polar of M is called the absolute bipolar of M. It is equal to the bipolar of the circled cover of M. (2) The polar M° of a set M cz Ex is convex, weakly closed and contains o. // M is circled, M° is absolutely convex. Indeed M° is the intersection of the weakly closed half-spaces $R(wx0):g 1 in E2, as x0 runs through all the elements of M, and so M° is weakly closed and convex; further the element o belongs to all these half-spaces. If M is circled, M° is the absolute polar of M, and so it is absolutely convex. Conversely, we have (3) M and the weakly closed convex cover C (M,o) of M and o have the same polar M°. For <R(wx)^l, for all ueM° and xeM, and also W(uo) = 0<l; further all the elements x of the real line segment joining xx and x2 satisfy $R(wx)fgl, if xx and x2 do. If M is a linear subspace, M° = M1, so that forming an orthogonal space is just a special case of forming a polar set. If C is a cone with vertex o, the polar set C° is again a cone with vertex o, which is given by the equation (4) <R(wx)^0 for all xeC. For if C contains x it contains px for all p>0, and so if $R(wx);g 1 we also have pR(ux)^l; this is only possible if $R(wx):gO. We now come to the theorem of bipolars: (5) The bipolar M°° of a subset M of Ex is the weakly closed convex cover C (o,M) of o and M. If M is convex and weakly closed, and if M contains o, then M = M°°. The absolute bipolar of a subset M of E± is the weakly closed absolutely convex cover \~{M) of M. If M is circled, M°° = C (M) is equal to the absolute bipolar of M. Proof. By (3) we may suppose that oeM, and that M is convex and weakly closed. By (l)c), Mc=M°°. If on the other hand x0$M, then by 7.(1) x0 and M are strictly separated by a closed real hyperplane. Since this does not pass through o, by 7. (2) it can be taken in the form $R(w0x) = l, where ${(u0y)i^\ for all yeM and 5R(w0x0)>l. But then x0 does not belong to M°°, and so M => M°°. The second part of (5) follows directly from the remarks we made about the relation between polars and absolute polars.
9. The polar of a neighbourhood of o 247 We denote by ^(E^ the collection of all convex weakly closed subsets of El which contain o. By 7.(6) it is enough to require the sets to be closed with respect to any topology compatible with the dual pair (E2,El}, instead of requiring them to be weakly closed. Under the relation CxaC2, C^) forms a complete lattice; /\Ca is the intersection p)Ca and \JCa= C Ca. a a a (6) Let <£2> £i> be a dual pair. The complete lattices ^(E^ and £(E2) are dually isomorphic, when each set is made to correspond to its polar. Proof. By (2), M°e(£(£2) (respectively ^(E,)) if Mett^) (respectively &(£2)), by (5) the correspondence is one-one, and by (l)b) it reverses the partial order, so that it is a dual lattice isomorphism (cf.§7,9.). In particular, because of the correspondence between /\ and \/ we have (7) // the sets Ca are weakly closed convex subsets of Ex containing o, we have (8) fK =cq. The dual formula holds for arbitrary subsets Ma of Ex: (9) ((jMaY=fVVC Since under the formation of polars absolutely convex sets, cones and linear subspaces correspond to absolutely convex sets, cones and linear subspaces, respectively, we obtain dually isomorphic lattices of weakly closed absolutely convex sets, of weakly closed convex cones and of weakly closed linear subspaces. This last of these results has already been obtained in 3. If the sets Ca are weakly closed absolutely convex sets or weakly closed linear subspaces Fa, respectively, (8) takes the form (io) (f]ca) =rc: or (ii) ((V.) =5X- respectively. 9. The polar of a neighbourhood of o. We begin by investigating the dual pair <£*,£>, where E is a vector space and £* is its algebraic dual. We have (cf. § 10,6.(3)) (1) // E is a vector space of dimension d over K, then E*[XS(E)] is topologically isomorphic to cod(K), the topological product of d copies of the field K.
248 §20. Duality Proof. If {xa} is an algebraic basis of £, each linear functional u on E is given by the corresponding coordinate vector u = {ua}, where va = uxa; conversely, every such coordinate vector u defines a linear functional on £, so that £* is algebraically isomorphic to a>d(K). In defining the weak neighbourhoods U of o it is sufficient to take finitely many elements xa of the basis, sup |wxa.|<6 means however that i = 1,..., n finitely many coordinates of the elements u of the neighbourhood U are bounded in modulus by c. The weak topology XS(E) therefore coincides with the product topology on cod(K). (2) // <£2>£i> is a dual pair, the completion of El[Xs(E2j] is £f [£s(£2)], and it is topologically isomorphic to cod(K), where d is the algebraic dimension of E2. Proof. Since £jc£* and E%[XS(E2)] is topologically isomorphic to cod(K)9 and further since a>d(K), being the topological product of complete spaces, is again complete, it is sufficient to show that E1 is dense in E% with respect to the weak topology ZS(E2). If this were not the case, the weak closure Ex of Ex in E\ would be different from E\\ then by 1.(3) there would be a non-zero weakly continuous linear functional v on £f which would vanish on Ex. This contradicts the fact that the weak dual of E\ is equal to £2, by 2.(3), and that there is no such non-zero v in £2. Using these results we can characterise the weakly bounded subsets of a locally convex space: (3) The weakly bounded subsets of a locally convex space E[%] are the same as the weakly precompact subsets of E. Proof. Every weakly precompact set is weakly bounded, by § 15, 6.(6). On the other hand, if Mis (Xs(Ef) bounded in £, M is also £S(F) bounded in the weak completion (£')* of £, and this is topologically isomorphic to some cod(K), by (2). But every bounded subset of cod(K) is relatively compact, by § 15, 6. We now prove the Alaoglu-Bourbaki theorem: (4) // U is a neighbourhood of o in the locally convex space £[£], U° is a weakly compact subset of E'. Proof. Suppose first that U is absolutely convex. To begin with we form the polar of U in E* => F. If the linear functional veE* belongs to this polar, |i?x|^l for all xeU. But this means that v is continuous, and is therefore an element of F. U therefore has the same polar in F as it has in £*. U° is weakly bounded in F and in E*. For given any xe£, U contains a suitable scalar multiple px, p>0, and it follows from this that given finitely many xt in £, sup \vxt\<M for a suitable M>0. i - 1,..., n veU°
10. A representation of locally convex spaces 249 By (2), E*[XS(E)] is topologically isomorphic to some cod(K), in which every bounded set is relatively compact. Thus U°, considered as a subset of £*, is relatively weakly compact; since it is weakly closed, by 8.(2), it is therefore weakly compact. Since every neighbourhood U contains an absolutely convex F, and since U° cz V°, U° is also weakly compact. We observe that there is a certain analogy between (4) and § 10,11.(2). In § 14, 5.(3) we introduced a norm on the dual space E of a normed space E. The closed unit ball in E is simply the polar of the unit ball in E. We therefore obtain as a special case of (4): (5) The closed unit ball of the dual of a normed space is weakly compact. As an application of (5) we prove: (6) The absolute bipolar of a sequence xn which converges weakly to o in the sequentially complete locally convex space £[£] is weakly compact, 00 00 and consists of all £ ^nxn with Z l£nl = l- n=1 n= 1 Proof. The xn form a weakly bounded subset N of £, by § 15,6.(4). By the theorem of Mackey, which will be proved in the next number but one (11.(7)), N is also ^-bounded in E. By 6.(1) the absolute bipolar m \~{N) is also bounded. All the partial sums £ £nx„ clearly belong to n= 1 \~{N\ and are ^-convergent, since £[£] is sequentially complete. 00 The collection M of all these £ ^nxn ls therefore a bounded subset of £[£] which lies in r(N). "=1 The mapping which sends each * = (£„) in Z1 to the element 00 Ax = Yj £>nxneE is a linear mapping from Z1 into E[%~\, underwhichM n= 1 is the image of the closed unit ball Kofi1.11 is the dual of c0, by § 14, 7. (11). (00 \ 00 Y,€nxn)= Z ^n(^^n); the sequence vn = (vxn) is con- n=l / n=l vergent to zero, and therefore represents an element v = (vn) of c0. It follows from the relation v(Ax) = vx which we have just established that A is a weakly continuous mapping for the topologies £s(c0) on Z1 and ZS(E) on E. The set K is weakly compact, by (5), and so therefore is its continuous image M. Consequently M is absolutely convex and weakly closed, so that it is equal to the absolute bipolar \~{N) of the sequence x„. 10. A representation of locally convex spaces. In §14,9. we introduced the (B)-space C{K) of all continuous real- (respectively complex-)
250 §20. Duality valued functions on the compact space K. The following representation of (B)-spaces follows from 9.(5): (1) Every (B)-space E is norm-isomorphic to a closed linear subspace of some suitable C(K\ where K is compact. Proof. We take for K the closed unit ball in the dual space £'ps(£)]. Each x js then a weakly continuous linear functional on K, and so is an element of C{K). The norm ||x|| is equal to sup|wx|, by § 17,6.(4), and ueK so it is equal to the norm of x as an element of C(K). In this way E becomes a closed linear subspace of C{K). In order to represent arbitrary locally convex spaces as function spaces we have to use locally compact spaces. Let R be a locally compact topological space. We denote by C(R) the vector space of all continuous real-(respectively complex-)valued functions on R. As topology X on C(R) we introduce the topology of uniform convergence on the compact subsets of R, which has as base of neighbourhoods of o the sets UK;E consisting of all f(x) with sup|/(x) | <£, where K is a compact subset of R. xeK The topology X is also defined by the semi-norms /?*;(/) = sup |/(x)|. xeK By §6,4.(5) there are "sufficiently many" continuous functions on R, i.e. for every two points x=t=y there is always a continuous function /with /(x) +f(y). C(R) is a locally convex space with respect to X. If R is compact, we obtain the special case considered at the beginning of this number. (2) C(R) is complete. IfR is countable at infinity, C(R) is an (F)-space. Proof. If 5 is a Cauchy filter of functions in C(R\ then for each x0eR we can form the corresponding Cauchy filter of values of the functions at x0. This has a limit /(x0). The function /(x) defined in this way is the uniform limit on each compact set K of functions in 5, and so it is continuous on each compact set K; it is therefore continuous on R, so that C{R) is complete. If R is countable at infinity, there is a sequence Kx cz K2 <= ••• of compact sets whose union is the whole of R, and with the property that every compact subset of R is contained in one of the sets Kn. But then the topology X is defined by countably many semi-norms pKn(x\ so that C(R) is metrizable; since it is complete, by the first part of the proof, C(R) is an (F)-space. As a generalisation of (1) we now obtain: (3) Every locally convex space E[X] is topologically isomorphic to a linear subspace of a suitable C(R\ where R is locally compact.
11. Bounded and strongly bounded sets in dual pairs 251 If, further, £[£] is metrizable, R can be taken to be countable at infinity. Proof. By § 18, 3.(7) it is sufficient to establish the result for a topological product F[X~\ = T\Ea of (B)-spaces Ea. Let Ka be the closed a unit ball of E'a equipped with the topology Zs(Ea), under which Ka is compact. The disjoint union R = [JKa becomes a locally compact a space when for each waeKa we take as base of neighbourhoods in R its neighbourhoods in the compact space Ka. Each compact subset of R is contained in the union of finitely many Ka. The topology on C{R) of uniform convergence on the compact sets is therefore given by the semi-norms pK(f) = sup\f(u)lK = Kaiv--vKan. If we consider each ueK element x = (xa) of F as an element of C(R\ by considering each xae£a as a continuous function on Ka, as in (1), then pK{x) coincides with the semi-norm p{x)= sup ||xa.|| on F[3f|; the assertion follows directly from this. i = i,...,n If £[£] is metrizable, F[X] is the product of countably many spaces £a, and R is countable at infinity. We shall obtain a stronger result than (1) for separable (B)-spaces in §21, 3. 11. Bounded and strongly bounded sets in dual pairs. We saw in 7. that the formation of the closed absolutely convex cover of a set M a E[X~\ depends only on the dual pair <£',£>, and that every compatible locally convex topology gives the same result. We shall now show that the bounded sets of a locally convex space E[Z~] are also the same for every topology compatible with the dual pair <£',£>, so that the concept of bounded set depends only on the dual pair. We shall obtain this result as a special case of a more general result. Starting from a dual pair <£2,£1>, the following definition of boundedness for a subset of Ex is particulary natural: A subset M of E{ is said to be £2-bounded if sup|wx|=/u(w)<oo for each ueE2. It can be seen at once that this is not a new concept: (1) The E2-bounded subsets of E± are the subsets of Ex which are bounded in the weak topology £s(£2)- The following definition does however produce a new concept: a subset M of Ex is said to be strongly £2-bounded if sup \ux\ = fi(B)<co for each Ex-bounded set B cz E2. ueB, xeM Clearly every strongly £2-bounded set is also £2-bounded. We shall see in §21,2. that strong boundedness also coincides with bound-
252 §20. Duality edness with respect to a suitable locally convex topology on E1. We now investigate the question of when a bounded set is strongly bounded. Let M be an absolutely convex weakly bounded subset of E1. The 00 linear span of M in £x is equal to E1M= \J nM. M is an absolutely n= 1 convex a-body in the space E1M, and its Minkowski functional defines a norm |jx||M on E1M. In this way ElM becomes a normed space. We observe that the norm topology on ElM is finer than the induced topology <XS{E2\ since each weak neighbourhood of o contains a scalar multiple ofM. We say that M is complete in itself if every sequence in M which is Cauchy with respect to the norm has a limit in M. If M is complete in itself then E1M is a (B)-space with M as closed unit ball. (2) Let M be an absolutely convex, bounded, closed, sequentially complete subset of the locally convex space E[%~\. Then EM is a(B)-space with closed unit ball M, and X induces a topology on EM which is coarser than the norm topology. Proof. X is coarser than the norm topology, since each 2-neigh- bourhood of o contains a scalar multiple of M. By hypothesis, M is ^-sequentially complete, and applying §18,4.(4) b) to EM, M is also sequentially complete with respect to the norm, so that EM is a (B)-space. We now obtain the Banach-Mackey theorem: (3) Let <£2>£i) be a dual pair. An absolutely convex weakly bounded subset of Ex which is complete in itself is strongly E2-bounded. Every absolutely convex, closed, bounded, sequentially complete subset M of a locally convex space E\%\ is strongly E'-bounded. We give two proofs of this important theorem. a) The first proof reduces the theorem to the theorem of Banach (§15,13.(2')). If B is an arbitrary weakly bounded subset of E2, the restrictions of the functional ueB to ElM form a set of continuous linear mappings from ElM into K, and this set is bounded on each xeElM, i.e. sup|wx|< oo. But then by §15,13.(2') sup |wx|<oo, ueB ueB,xeM so that M is strongly £2-bounded. This establishes the first part of (3). The second part follows from the first, and from (2), since M is complete in itself, by (2), and is weakly bounded. b) The second proof uses the "sliding hump" technique of H. Lebes- gue and O. Toeplitz (cf. F. Hausdorff [1], for example). We first consider the case where M is complete in itself, and suppose that M is not strongly £2-bounded.
11. Bounded and strongly bounded sets in dual pairs 253 There then exists a weakly bounded set B cz E2 and a sequence uneB with v(w„)=:sup|wnx|->oo. On the other hand /i(x) = sup|wx| <oo for each xeEi. For each un there exists an xneM with (4) \unxn\^±v(un). We now determine a sequence nx <n2<'" of integers for which (5) ^v(uj^-ii(xj+ -2fi(xn2) + '" + ^Ifi(xnk_l) + k = Rk + k. Since ||x„J:gl, and since M is absolutely convex and complete in 00 J itself, the sequence £ -^ x„k is convergent in £1M to an element x0eM. We decompose x0 into three parts: (6) i ... J_ \ 1 (_}_ ^ + 7**"* + 4 Then for each wgB we have and uyk\ ^ - MxJ +' • • + -^zj v(Xnk-,) = Rk \uzk\ ^ -^ v(u) + -^-2 v(u) + •■• = — v(w). Thus \ux0\ ^ ^ |wx„J - Rk - ^ v(w). In particular, writing u = u„k, we obtain 2?VW"^VW"S kk*0| ^ ^ V(l/„k) - -~k V(UJ - —r V(uj + k = k, because of (4) and (5); this contradicts the fact that B is bounded. The proof also holds if we make the hypothesis on M that M is absolutely convex and bounded, and that every 2-Cauchy sequence in M has a limit in M, where X is a locally convex topology on Ex which induces a coarser topology on ElM than the norm ||x||M. For CO | Z ~Akxnu *s tnen a^so ^-convergent to an x0eM. n = l 4
254 § 21. The different topologies on a locally convex space The second part of (3) follows from this remark, since on any bounded subset M of a locally convex space E[X~\ the topology X is coarser than the norm topology of E1M. As a first corollary of (3) we derive the theorem of Mackey which was announced at the beginning of this number: (7) Suppose that <£2>£i) *s a dual pair. All the compatible locally convex topologies on E{ define the same collection of bounded sets. In particular the bounded and weakly bounded subsets of a locally convex space £[£] are the same. Proof. It is sufficient to prove the second assertion, and so it is sufficient to show that a weakly bounded subset M of E[X~\ is bounded in the topology X. If V is a closed absolutely convex ^-neighbourhood of o, U=U°°, by the theorem of bipolars, and U° is weakly compact in F, by 9.(4). The second part of (3), applied to U° a E[XS(E)], shows that U° is strongly E-bounded, i.e. sup \ux\ = fi<co. This means ueU ,xeM that Mc=iuC/°° = /uC/, i. e. that M is ^-bounded. The next result is a special form of the Banach-Mackey theorem which is convenient for applications: (8) // the locally convex space E[X~\ is sequentially complete, every weakly bounded subset of E (respectively of E') is strongly E'-bounded (respectively E-bounded). For by (3) all the absolutely convex bounded closed subsets of E[X~\ are strongly bounded, and, by (7), so are all the weakly bounded subsets B of E. But this means that if B' is a weakly bounded subset of E sup \ux\ < oo holds for each B, i. e. B' is strongly bounded. ueB',xeB We end with an example. In the dual pair <<p,<p>, where <p is the space of 00 finite coordinate-vectors 3E = {<J1,...,<J„,0,0,...}, with bilinear form ux= £ Vi€i9 a subset M c: cp is (^-bounded if and only if there are numbers mf > 0 with \Zi\Smi for £ = 1,2 for all xeM. In contrast, M is strongly bounded if and only if in addition there exists an n0 such that £„ = 0 for all seM, n>n0. § 21. The different topologies on a locally convex space 1. The topology Xm of uniform convergence on 9W. In §20,2., given a dual pair <£2>£i>> we introduced the weak topology XS(E2) on the space El. The definition given there can also be formulated in the following way:
1. The topology 1^ of uniform convergence on 5R 255 We consider the class g of all finite subsets F of E2\ the absolute polars (f~ F)° of the sets F form a base of neighbourhoods of o for XS(E2); alternatively the semi-norms pF(x) = sup\ux\ define the topology %(E2). It is natural to try to find other classes $ft of subsets of E2 which can be used to introduce locally convex topologies on Ex. If M is a subset of E2, we set pM(x) = sup\ux\. Using this terminology, we have ueM (1) If M a E2, pM(x) is a semi-norm on E1 if and only if M is weakly bounded in E2. For if M is bounded in E29 Pm(x)<co f°r eacn xeEu so that pM(x) is a semi-norm. On the other hand if pM(x) is a semi-norm, sup|wx|<oo for each x, so that M is bounded in E2. ueM Right from the beginning, then, the only suitable classes are classes $ft of bounded sets in E2. We call such a class $ft total in E2 if its union (J M is total in E2 for the topology ^(i^). Mean (2) Suppose that 9M = {M} is a collection of bounded subsets of E2. The topology XOT defined by the system of semi-norms {pM(x)} on hi is a locally convex topology on E{ if and only if $ft is total in E2. Proof. By § 18, 1.(3), %m is a locally convex topology on El if and only if XOT is Hausdorff; this happens if and only if the linear sub- space N of all those xeE1 with pM(x) = ® f°r a^ Me9M consists ofo alone. But if pM(x) = 0 for all Me9M then xe (] M1 = (|jM)-L, and con- Mean versely for such an x pM(x) = 0 for all M. But 9M is total in E2 if and only if ( (J mY=o. Vean ' The locally convex topology H^ is also called the topology of uniform convergence on the sets M of 9M, or on 9M. Here, the xeEx are considered as functions, namely as linear functional on E2. A base of XOT-neighbourhoods of o is obtained when we form the absolute polars {\~ M)° of the sets M of 9M, form their scalar multiples p([~ M)°, and then form the finite intersections of these. When do two topologies XOTl and X^2 coincide? A class 9M={M} is said to be saturated if the following conditions hold: 1) If M belongs to 9M, so does every subset of M. 2) If M belongs to 9M, so does every scalar multiple pM, peK. 3) If Ml and M2 belong to 9M, so does their weakly closed absolutely convex cover p(Ml9M2).
256 §21. The different topologies on a locally convex space If 9M is an arbitrary class of bounded subsets of E2, we can form the smallest saturated class 9Jt which contains 9M. This saturated cover of 9M again clearly consists only of bounded subsets of E2. A little earlier we gave a base of ^-neighbourhoods of o. If SCR is saturated, we have: (3) // 9M is total and saturated, the polars M° of the absolutely convex weakly closed Me9M form a base of ^^-neighbourhoods of o in E1. For if M° is the polar of an weakly closed absolutely convex set in 9W, so is (-M) = pM°, and if M\ and M°2 are, so is r(M1,M2)° The next result answers the question we raised above: (4) The topologies %m and %& on Ex are the same; two topologies 3OTl and 3OT2 coincide if and only if 9Jl1=3Jl2- // 90?! cz35l2, %$li is coarser than 3gft2, and the converse also holds. Proof, a) In order to show that 3^ = 3^ it is sufficient to prove that the absolute polars of sets which are obtained from the sets of 9M by the processes of 1), 2) and 3) are again 3<m-neighbourhoods of o. If Me9M and JV c M, (r N)° => (rM)°, so that {r N)° is a 3<m-neighbourhood of o. It follows from {\~ pMf = - {\~ M)° that P (\~p M)° is also a 3OT-neighbourhood of o. Finally \~(M1,M2)° = (r Mxf n(r M2)° is a 3OT-neighbourhood of o. b) Suppose that 91 and 9M are saturated, and that 3^ = 3^. It is sufficient to show that Wcz9M. If JVeSR, {r N)° is a 3OT-neighbour- hood of o, and so by (3) there is an Me9M with M°° = M and M°cz(riV)0. It follows from this that M = M°° => (FN)00 => N. N, being a subset of M, is therefore also in 9M, so that 91 c 9M. We remark that the class g of finite subsets of E2 which we introduced at the beginning is not saturated. The saturated cover § clearly consists of all the bounded finite-dimensional subsets of E2. 2. The strong topology. Among the locally convex topologies Xm on El which can be defined in terms of the dual pair <£2,£1> there is a finest one, by 1.(1), namely the topology of uniform convergence on all the weakly bounded subsets of E2. For brevity, this is called the strong topology %h{E2) on E1. By 1.(3), the collection of all sets B°9 where B is a weakly closed bounded absolutely convex subset of E2, forms a base of neighbourhoods of o for the strong topology. It is also given by the system of semi-norms pB(x) = sup\ux\.
2. The strong topology 257 A subset M of E1 is bounded with respect to the strong topology if it is strongly bounded in the sense of the definition of § 20, No. 11. The Banach-Mackey theorem gives a sufficient condition for the bounded subsets of E1 to coincide with the strongly bounded sets. This does not always happen, as was shown by an example given in §20,11. Since, on the other hand, every locally convex topology compatible with the dual pair defines the same bounded sets as the weak topology, by Mackey's theorem (§ 20,11.), there are situations in which the strong topology on £2 is not compatible, and so has a larger dual space than E2. We shall consider this question in more detail in § 23. If E[X] is a locally convex space, we already have an abundance of possible ways of introducing a locally convex topology on the dual space E'. If SCR is any total class of bounded subsets of E with respect to either X or XS(E) (cf. §20, 3.(4)), £'[XOT] is a locally convex space. In particular we call the space E' with the strong topology Xh(E) the strong dual of E[X~\. In the case of a normed space £, the strong dual has already been introduced in § 14, 5., for the norm introduced there on E clearly defines the strong topology on E. We now give two characterisations of the strong topology. Following the terminology of Bourbaki, we call a closed absorbent absolutely convex subset of a locally convex space E[X] a barrel. By §21,7.(6), we can require the sets to be closed in any compatible topology, e. g. the weak topology. (1) Suppose that <£2,£1> is a dual pair. The barrels in E1 form a base of neighbourhoods ofo for the strong topology Xb(E2) on Ex. Proof. If U is a barrel in El9 U°, being the polar of an absorbent set, is weakly bounded. By the theorem of bipolars, U = (U°)°, so that U is a strong neighbourhood of o. Conversely every strong neighbourhood B° of o, where B is weakly bounded and absolutely convex in £2, is a barrel in E1. Following Bourbaki, a locally convex space E[X] is said to be barrelled if the barrels form a base of X-neighbourhoods of o. From (1), we get: (2) The barrelled spaces E [X~\ are those locally convex spaces whose topology X coincides with the strong topology Xh(E). The next result gives the second characterisation of the strong topology: (3) // E\X] is locally convex, the semi-norms p(x) which are lower semi-continuous with respect to X coincide with the Xb(E)-continuous semi-norms. 17 Kothe, Topological Vector Spaces I
258 §21. The different topologies on a locally convex space X is therefore the strong topology if and only if every lower semi-continuous semi-norm on E [2] is continuous. Proof. By §6,2.(3) the lower semi-continuity of a semi-norm p(x) implies that the subsets of E[%] defined by p{x)^y are closed. These sets are therefore closed absolutely convex a-bodies, i.e. barrels. Similarly, by §6,2.(3) the semi-norm corresponding to a barrel is lower semi- continuous. The assertion therefore follows from (1). 3. The original topology of a locally convex space; separability. We have just given a condition for the original topology % of a locally convex space E[X~\ to be the strong topology. We shall now show that in every case X is one of the topologies Xm defined in 1. We recall the concept of an equicontinuous set of mappings (§ 15,13.). The elements u of E\ being continuous linear functionals, are continuous mappings from E[%] into K. Considering them in this way, we have: (1) A subset M ofE is %- equicontinuous if and only if M c UG, where U is a suitable absolutely convex neighbourhood ofo in E[%]. For by § 15,13.(1) the equicontinuity of M is equivalent to the existence of an absolutely convex X-neighbourhood (/ of o with sup |wx| = l, i.e. M c U°. ueM, xeU (2) Suppose that E[X] is locally convex. Let © denote the class of all %-equicontinuous subsets ofE'. © is total and saturated in E\ and <X = <£®; 2 is therefore equal to the topology of uniform convergence on the %-equi- continuous subsets ofE'. For every locally convex topology %m on E, the saturated cover 9M is the class of all %m-equicontinuous subsets of E'. Proof. The closed absolutely convex ^-neighbourhoods ofo form a base of neighbourhoods ofo. Their polars U° are equicontinuous, and, since U=U°°, X is the topology of uniform convergence on the sets U°, so that 2 = 2«>. © is saturated since pUc=(-Uj and r(U°1,U°2) = (U1nU2)G again belong to ©. \P J The second assertion follows from 1.(4) and the definition of equicontinuity. In the special case of normed spaces, this result has already been proved in §17, 6.(3). If <£2,£1> is a dual pair, every compatible locally convex topology is therefore a topology Xm. We saw in 2. that the strong topology need not be compatible, and it is finer than every compatible locally convex topology.
3. The original topology of a locally convex space; separability 259 Conversely we know (§20,2.(4)) that the weak topology ZS(E2) is the coarsest compatible locally convex topology on E^ A locally convex topology on E1 which is strictly coarser than the weak topology will therefore not be compatible, and will produce a smaller dual than E2. We now show that this can happen. Let N be a weakly total subset of E2. We denote by ZS(N) the topology of uniform convergence on all the finite subsets of N. By 1.(2), ZS(N) is locally convex on El9 and it is coarser than ZS(E2). In particular if N is a weakly dense linear subspace H^E2 of E29 then XsiH) is simply the weak topology on E1 defined by the dual pair <//,#!> (which is a dual pair, since H is dense in E2). By § 20, 2.(3), the dual of El\%s{H)'\ is H> an<^ so ^ *s smaller than E2. The topology £s(//) is therefore not compatible with the dual pair <E2>£i>> and it is coarser than every compatible topology. This topology can however coincide with the weak topology on certain subsets of Ex: (3) Suppose that E[%~\ is locally convex and that M is a %-equicontin- uous subset ofE'. Then ifN is a weakly total subset of E[%] the topologies XS(E) and ZS(N) coincide on M. Proof. We can take M=U°, by (1). By §20,9.(4), U° is weakly compact. The locally convex, and therefore Hausdorff, topology ZS(N) on E is coarser than the weak topology; it must therefore coincide with it, by §3, 2.(6). From (3) we obtain (4) If the locally convex space E[%] contains a countable total subset N, every equicontinuous subset M of E' is metrizable in the weak topology, and is weakly sequentially separable. For ZS(N) has a countable base of neighbourhoods of o in E, so that E' is metrizable under XS(N), and, by (3), M is metrizable under the weak topology. Since further M is weakly relatively compact, by § 4, 5.(2) M is weakly separable in the sense that every element of M is limit of a subsequence of some fixed sequence of elements of M. (5) // £[X] is a separable metrizable locally convex space, E' is weakly sequentially separable. For if the absolutely convex sets Un,n=l,2,..., form a countable 00 base of neighbourhoods of o in E[%~\, E'= \J U°n, and each U° is weakly sequentially separable, by (4). n=1 Consequently the assertion also holds for E. As an application of (5) we prove a result of Banach and Mazur relating to §20,10.(1) (cf. Banach [3]): 17*
260 §21. The different topologies on a locally convex space (6) Every separable normed space E is norm-isomorphic to a linear subspace of C(7), the space of continuous functions on the interval / = [0,1]. Proof. The closed unit ball K of E is a compactum under the weak topology, as in the proof of (5). By a well-known theorem (cf. Hausdorff [2], p. 132 or Bourbaki [5], Vol.4, p. 31, ex.18) K is the continuous image of Cantor's ternary set JC0 cz/, which is nowhere dense in /. If the point teK0 there corresponds the continuous linear functional uteK, we set x(t) = ut(x). x(s) is defined on the disjoint components of the open set /~K0 by linear interpolation. This function x(s), defined on the whole of J, is continous on J: it is sufficient to prove this on K0. But if x is a weakly continuous linear functional on K, then, because of the continuity of the mapping from K0 onto K, x also defines a continuous function on K0. The correspondence x->x(s) is clearly linear, and it is also an isometry, since ||x|| = sup|wfx| = max|x(s)|. ut se/ 4. The Mackey topology. We shall now give a precise characterization of the topologies XOT which are compatible with a dual pair <£',£>. By §20,9.(4) the polar U° of an absolutely convex neighbourhood U of o of a locally convex space E[%] is absolutely convex and weakly compact in E'. The class © of all X-equicontinuous subsets of E is therefore contained in the class ft of all absolutely convex weakly compact subsets of E'9 together with their subsets. Since all the sets of ft are bounded in F, the topology Zk(E') of uniform convergence on all the sets of ft is a locally convex topology on E which is finer than the original topology. We call Zk(E') the Mackey topology on E. Like the strong topology, it depends only on the dual pair <F,£>, and not on the original topology. (1) The class ft of all absolutely convex weakly compact subsets of E\ together with their subsets, is total and saturated, and therefore consists of all the %k(E')-equicontinuous subsets of F. A base of neighbourhoods of o for the Mackey topology is formed by the sets K°9 as K runs through the absolutely convex weakly compact subsets ofE'. The first part follows from the fact that if the sets Kt are weakly n compact and absolutely convex then V Kt is also weakly compact and i=l absolutely convex, by §20,6.(5). The second part follows from 1.(3). Because of the counterexample in § 20, 6., the class ft' of all weakly relatively compact subsets of E' is in general larger than ft and the corresponding topology Xfi. is therefore different from the Mackey topology. We now prove theMACKEY-ARENS theorem (Mackey [5], Arens [1]):
4. The Mackey topology 261 (2) Suppose that <£2»^i) *s a dual pair. The Mackey topology Zk(E2) is the finest locally convex topology on E1 which has E2 as dual space, and so it is the finest locally convex topology compatible with the dual pair <£2^i)« Proof. We have already shown that every compatible locally convex topology on E1 is coarser than the Mackey topology. It remains to show that E2 is the dual space of E1[Xk(E2)]. We consider the dual pair {E^E^, where E\ is the algebraic dual of E1. Let u0 be a Xfc-continuous linear functional on E1. It lies in Ef, and by (1) there is an absolutely convex weakly compact subset K of E2 for which sup|w0x|gl; u0 therefore belongs to the polar (K°)° of K° in xeK° Ef. Since E2 c E\, K is also absolutely convex and weakly compact in EX, and so it is weakly closed. By the theorem of bipolars, applied to the dual pair {E^E^, we therefore have K = K°°. Thus it follows from u0eK°° that u0eKczE2. The dual space of E1[Zk(E2)] is therefore a subspace of E2. Since conversely every element of E2 defines a Zk- continuous linear functional on El9 (2) is established. From (2) and § 20, 2.(4) there follows directly (3) // E[X] is a locally convex space, the original topology % is finer than the weak topology ZS(E') and coarser than the Mackey topology Zk(E'). The Mackey topology in turn is coarser than the strong topology %{E'). All the locally convex topologies on E1 which are compatible with the dual pair (E2,Exy are obtained by forming the topologies X^, where $Jl is a class of subsets of E2 with g cz 9M c ft, where g is the class of finite subsets oj'E2 and ft is the class defined above. We remark that analogous results were obtained for linearly topologized spaces in §10; in particular §10,11.(4) is the analogue of the Mackey-Arens theorem. (4) // £[X] is barrelled, % coincides with both the strong and the Mackey topologies, and every bounded subset of E' is weakly relatively compact. For by 2. (2) X is the strong topology, and so it is finer than the Mackey topology. On the other hand £ is compatible with the dual pair {E',E}, and so X must be coarser than the Mackey topology. The identity of the topologies implies the identity of the equicontinuous sets, so that every bounded subset of E' is weakly relatively compact; conversely a weakly relatively compact set is always bounded. (5) The completion £[X] of a locally convex space E[%] has the same dual space E' and the same equicontinuous sets in E'.
262 §21. The different topologies on a locally convex space The topologies ZS(E) and Xs(£) coincide on the equicontinuous sets of the dual of a locally convex space E[X]. lf% is the Mackey topology of E\%\ X is the Mackey topology of £[£]. Proof. The identity of the duals was proved in § 15,9.(11). If U is a closed absolutely convex neighbourhood of o in E[%], the closure U~ in E has the same polar U° = U° in £', and so the X-equicontin- uous and X-equicontinuous subsets coincide. By §20,9.(4), the polar U° of a neighbourhood U of o in E[%] js Xs(£)-compact, and from what we have just proved it is also ZS(E)- compact; since Xs(£) is finer than ZS(E)9 the two topologies coincide on U°, by §3,2.(6). Applying this to the Mackey topology X, the classes of absolutely convex Xs(£)-compact sets and of absolutely convex Xs(£)-compact sets coincide. But using the first result of (5) this means that % is the Mackey topology on E. In analogy to § 10,12.(2), we have: (6) Suppose that £i[X] and E2\%2\ are locally convex. A linear mapping from El into E2 is %k-continuous if and only if it is weakly continuous. Every continuous mapping is %k-continuous. Proof. By § 20,4.(5) every Xfc-continuous mapping is weakly continuous. Suppose conversely that A is ^-continuous. The mapping A' is weakly continuous, by § 20, 4. (6), and it therefore maps every absolutely convex weakly compact subset K c E'2 into an absolutely convex weakly compact subset A'(K)czE\. But then it follows from v(Ax) = (A'v)x thattheimage A(A'(K)°) of the ^-neighbourhood A\K)° c El lies in the given ^-neighbourhood K° c E2; A is therefore ^-continuous. Example. It is easy to see that for the dual pair <<p,<p> considered at the end of §20,11. Xk((p) = Xs((p)> 5. The topology of a metrizable space. Most of the locally convex spaces met with in applications have a natural original topology X, which coincides either with the strong or the Mackey topology. Here we shall establish this fact for metrizable spaces. (1) Suppose that E[%] is locally convex. Every absolutely convex weakly compact subset ofE' is strongly bounded. This is a special case of the Banach-Mackey theorem (§20,11.). If we call the topology of uniform convergence on the str£ngly bounded sets of E2 the topology (Xb*(E2)9 we can formulate (1) in the following way: (2) The topology %b*(E2) is always finer than the Mackey topology
6. The topology Xc of precompact convergence 263 For metrizable spaces we have (3) If the locally convex space E[%] is metrizable, % is equal to the Mackey topology %k{E') and to the topology %h*(E'). If, further, E[%] is complete, so that it is an (F)-space, % is equal to the strong topology Xb(E'); thus all (F)-spaces are barrelled. Proof, a) Suppose that £[£] is metrizable. By the Mackey-Arens theorem, X is coarser than Zk(E')9 and Zk(E') is coarser than Zh*(E')9 by (1). The first part of (3) will therefore be established if we can show that every strongly bounded subset B of E' is £-equicontinuous, and so if we can show that there is a X-neighbourhood U of o with B c U°. The topology X is defined by an increasing sequence of semi-norms \\x\\n on E. We suppose that B is not X-equicontinuous. Then for each n sup \ux\ = oo. ueB,\\x\\n£l There is therefore a sequence uneB, xneE, \\x„\\n^l for which \unxn\>n for n=\,2,.... It follows from sup ||xjk^max(||x1||k,...,||xjk, l)<oo w = l,2,... that the xn form a bounded subset X of E; but then the strong bound- edness of B implies that sup |wx„|<oo, which contradicts \unxn\>n. ueB,xneX b) If E[%] is complete, the strongly bounded and bounded sets in E' coincide, by §20,11.(8), so that the two topologies Zb(E') and %b*{E') are the same. The second part of (3) can also be expressed in the following way: (4) In the dual E' of an (F)-space E[X] the following classes of sets coincide: a) the bounded sets, b) the strongly bounded sets, c) the weakly relatively compact sets and d) the equicontinuous sets. We now give an example of a normed space E on which the strong and the Mackey topologies are different. By (3) this is the case if there is a bounded set in E which is not strongly bounded. As in §20,6., let E[i] be the space (p equipped with the /2-norm topology. Then E = l2, and the strongly bounded sets of E are the /2-bounded sets, while the weakly bounded sets of E are the subsets whose elements x = (£„) satisfy inequalities \£„\^M„, ft =1,2,..., for arbitrary non-zero M„. 6. The topology Xc of precompact convergence. If E[Z] is locally convex, the X-precompact subsets of E form a class of bounded sets which cover E, and so the topology Xc of uniform convergence on the precompact subsets of E[%~\ is a locally convex topology on the dual space E'. In contrast to the topologies Zh(E)9 Zk(E) and £/,*(£), this topology depends upon the original topology, and not just upon the dual pair <£',£>.
264 §21. The different topologies on a locally convex space By § 20, 6., the class (£ of all precompact sets in £[X] is saturated; by contrast this need not be the case for the class of all relatively compact subsets of E, since the closed absolutely convex cover of a compact set need not be compact. (1) // E[%] is quasi-complete, the topology ZC(E) on E' is coarser than the Mackey topology Zk(E)9 so that (E'[Xc(Ey\)' is again equal to E. Proof. If E[X] is quasi-complete, every absolutely convex precompact subset of E is relatively compact. Every X-compact set C is also weakly compact, by § 3, 2.(6), so that (£ c ft, where ft is the class defined in 4. It follows from (£ cz ft that Xk is finer than ZC9 so that, applying the Mackey-Arens theorem, E is again the dual of F[3J. The following theorem gives an important property of the topology Zc: (2) Suppose that E[%] is locally convex. On every equicontinuous set M cz E the topologies ZS(E)9 %(E), ZS(E) and ZC(E) coincide, where E[i~\ is the completion of E[%]. Proof, a) Suppose that M is equicontinuous and that w0eM. In order to establish the identity of ZC(E) and £s(£) on M it is sufficient to show that, given a precompact set C cz E, we can find finitely many x(eE with the property that, if ueM and sup \(u — u0)xi\< 1, then sup|(w-w0)y|<2. i = i,...," yeC Because of the equicontinuity of M, and therefore of M — u0, there is a X-neighbourhood U of o with sup \(u — u0)z\<\. C is totally ueM,zeU n bounded, so that there are finitely many x{eC with C cz \J (Xi+U). i= 1 Each yeC therefore takes the form y = xt + z, zeU. For all ueM with sup|(w — M0)xf|<l we then have sup|(w — u0)y\ S sup \(u — u0)Xi\ + sup|(w — u0)z\ <2. yeC i = l,...,n zeU b) By 4.(5), E[%~\ has the same equicontinuous sets as E[X], and ZsiE) and ZS(E) coincide on the equicontinuous sets. The assertion about ZC(E) follows from this. The following strengthening of the Alaoglu-Bourbaki theorem follows from (2): (3) Every equicontinuous set of the dual E' of a locally convex space £[3f] is relatively compact with respect to the topologies %C{E) and If U is a X-neighbourhood of o, U° is therefore compact with respect to these topologies. Compare 3.(3), as well!
6. The topology Xc of precompact convergence 265 We know that the strong dual of a normed space is complete (§14, 5. (5)). A more general and stronger result is (4) // E[X~\ is metrizable and locally convex, E is complete under both the strong topology and the topology £C(E). // £[£] is an (F)-space, E is also complete under the Mackey topology Xk{E). Before proving this we establish a lemma due to Grothendieck : (5) A linear mapping A from a locally convex space E[%~\ into a locally convex space F[jX'] is uniformly continuous on an absolutely convex subset M cz E if and only if it is continuous at o on M. Proof. Suppose that A is continuous at o on M; then given an absolutely convex X'-neighbourhood K of o there is an absolutely convex ^-neighbourhood U of o with i(Mn[/)c|K If x and y are two elements of M with x — yeU, then x — yelM, since M is absolutely convex. From this it follows that x — ye(2M)nU = 2(Mn^U). Consequently ——-eMnU, so that Al—-—\e-V,A(x — y)eV. This means that A(x — y)eV if x — yeU, so that A is uniformly continuous on M. In particular a linear functional is ^-continuous on M if it is ^-continuous at o on M. Proof of (4). a) First we show that E[%c(Ej] is complete. A 2C-Cauchy filter g={Fa} on E is certainly a 2S(£)-Cauchy filter, and so by §20,9.(2) it has a ^-adherent point w0e£*. If r,>0 and C is a precompact subset of £, then sup|(w' — u)x\ ^ for all u,u' in a suitable ¥\ and so xeC 2 (6) sup|(w0 — u)x\ ^- for t/eFa. xeC 2 u0 is therefore also a ^-adherent point of g. We now show that u0 is ^-continuous on every absolutely convex precompact set, and so on every precompact set. Each u is ^-continuous at o in C, and so for any one u in Fa there £ is a ^-neighbourhood U of o with sup \ux\^-; (6) then shows that Jcel/nC 2 sup |w0x|^g, so that uQ is ^-continuous at o in C. It then follows xeUnC from (5) that u0 is ^-continuous on C. If u0 were not continuous on E, then since E[%] is metrizable, there would be a sequence xneE converging to zero with uo{xn)-/>0. But the subset C0 of E consisting of o and the xn is compact, and u0 would not be continuous on this set C0, contradicting what we have just proved.
266 §21. The different topologies on a locally convex space Consequently u0 belongs to £', and F[jXc(£)] is complete. b) The strong topology Xh(E) is finer than XC(E) on E'. The polars B° of the bounded subsets B of E form a base of neighbourhoods of o for the strong topology, and the sets B° are weakly closed, and a fortiori £c(£)-closed, so that by § 18, 4.(4) b) E' is strongly complete as well. If E[X~\ is an (F)-space, the topology Xk(E) on E' is finer than XC(E), by, (1), and so E' is also ^-complete. 7. Polar topologies. We have the following theorem of Grothen- dieck: (1) Suppose that <£2>£i> *s a dual pair, that 9JI is a saturated collection of bounded subsets M of Ex which cover Eu and that 91 is a similar collection in E2. The four following statements are equivalent: a) Each MeWl is X^-precompact. P) Each Ne9l is Xm-precompact. y) The topologies X^ and Xs coincide on each MeSR. S) The topologies Xm and Xs coincide on each Ne9l. Proof. The sets Ne$l are 3^-equicontinuous subsets in (Eipgj])' => E2. By 6.(2) the topology Xs coincides with the topology of ^-precompact convergence on N. If a) holds, X^ is coarser than this last topology, and, since [j M = EU it is finer than X^E^; X^ and MeSPt Xs therefore coincide on N. Thus 5) follows from a). We can however derive /?) from S), for the 3^-equicontinuous subsets N are weakly bounded in (Ei[£OT])' and are therefore weakly precompact in E2 by §20, 9.(3); they are therefore also ^-precompact, since Xs and Xm are the same on the sets N. Because of the symmetry of the situation, y) now follows from /?), and a) from y). As before let 91 be a saturated collection of bounded sets covering E2, and let X be the topology X^ on Ex. We call the topology ^(^[j!]) on E2 the topology X° polar to X. From (1) there follows (2) a) The topology X° polar to X = Xm is the finest locally convex topology of the form Xm(Ei) on ^2 which coincides with X^E^ on all the X-equicontinuous subsets N of E2. b) The topology X°° bipolar to X on Ex is always finer than X, and it is equal to X if and only if the collection of all the X-equicontinuous subsets of E2 consists of all the X°-precompact sets. c) // 3^ is finer than X2 on Eu then X\ is coarser than X°2. d) X000 = X°.
8. The topologies 27 and Zlf 267 Proof, a) 2° is a topology which coincides with 2S(£1) on the 2-equicontinuous subsets Ne9l. On the other hand if 2' = 2OT is a second such topology on E2, then S) holds. Theorem (1) then implies that a) holds; i.e. each Me^R is 2-precompact, and so 2OT is coarser than 2°. In particular, putting 2^ = 2°, the class of 2°-equicontinuous subsets of E1 consists of the collection of all 2-precompact subsets of Ex. b) The collection 91 of all the 2-equicontinuous subsets of E2 consists of 2°-precompact sets, while the collection 91' of all the 2°°-equicontinuous sets consists of all the 2°-precompact sets. It therefore follows from 9t <= 9t' that 2°° is finer that 2. c) This assertion is immediately obvious. d) On the one hand 2000 = (20)00 is finer than 2°, by b), and on the other 2°° is finer than 2, again by b), so that 2000 = (200)° is coarser than 2°, by c). (3) Suppose that <£2>£i) *s a dual pair. The topology polar to the weak topology 2S(£2) is the strong topology 2b(£1), and the topology polar to this is the topology of strongly precompact convergence. For since the weakly precompact subsets of Ex coincide with the bounded sets, 2S(£2)° is equal to 2b(£1). The topology bipolar to the weak topology is therefore different from the weak topology, in general; in contrast the strong topology, being the topology 2S(£2)°, is its own bipolar, by (2)d). 8. The topologies %f and 27/. Suppose that (E2,Exy is a dual pair, and that 2 is a topology <X9l{E2) on Ex lying between 2S(E2) and Zb{E2). We saw in 1. that 2° = 2C(£1[2]) is the finest topology 2aR(£1) on E2, defined by a saturated class of bounded subsets of Ex covering Eu which coincides with the weak topology 2s(Et) on the 2-equi- continuous subsets of E2. We now denote by %J (respectively %lf) the finest topology (respectively the finest locally convex topology) on E2 which coincides with 2S(£1) on the 2-equicontinuous subsets of E2. As we shall see later, 2/ is generally different from %lf, and so need not be locally convex. However, we have: (1) The topology Xf on E2 is Hausdorff and translation-invariant, and it has a base of circled absorbent neighbourhoods of o. Proof. The topology 2/ is determined by giving the collection of 2/-closed subsets of E2. Let 51 be the collection of all subsets A of E2 whose intersections AnM with all the 2-equicontinuous subsets M of E2 are weakly closed in M. It is easy to see that this collection 51 satisfies the axioms for the closed sets of a topology 20 (§1,1.)- ^o
268 §21. The different topologies on a locally convex space coincides with £s(£i) on the sets M, since it gives the same closed sets in M as the weak topology does. On the other hand A is the most extensive class of subsets of E2 for which this is the case, and so X0 is equal to 3/. The open ^/-neighbourhoods of o are the sets U whose intersections UnM with the 3>equicontinuous sets M containing o are weak open neighbourhoods of o in M. But then the sets x0+U are the open ^/-neighbourhoods of x0, for x0 + M is 3>equicontinuous if M is, and so (x0 + U)n(x0-\-M) = x0-\- UnM is a weak open x0-neighbourhood in x0 + M. The topology 3/ is therefore translation- invariant. The open ^/-neighbourhoods U of o form a filter base, and the intersection of all the sets U reduces to the point o, since this is already so for the weak neighbourhoods of o. Each U is absorbent, for every ueE2 belongs to some absolutely convex 3>equicontinuous set M, and so UnM contains a suitable scalar multiple of w. Finally each U contains a circled open ^/-neighbourhood of o. To prove this, let V be the set of all those x for which the set X of all ax, |a| ^ 1, lies in U. If M is an absolutely convex 3>equicontinuous set, VnM is a weak neighbourhood of o in M contained in UnM. We show that VnM is weakly open in M, which implies that V is a circled 3/-open neighbourhood of o. If xeVnM, X is weakly compact in the open subset UnM of M. By § 15, 6.(9) there is an absolutely convex weak neighbourhood W of o in E2 with (X+W)nM <=UnM. But then if ye{x-\-W)nM, o>y lies in UnM for every |a|^l, so that yeVnM; VnM is therefore weakly open in M. Komura [2] has given an example to show that E2 [3/] need not be a topological vector space. A base of neighbourhoods of o for the topology %lf is clearly obtained by taking all the absolutely convex ^/-neighbourhoods of o; a base is also obtained by taking all the convex ^/-neighbourhoods of o, for by (1) every ^/-neighbourhood of o contains a circled ^-neighbourhood of o. We now give an example due to Klee [2] III to show that 3/ and %lf can be different. (2) Suppose that El is the space cod, that E2 is cpd and that X is the weak topology 2s(<pd) on cod. 3/ and Xlf coincide on E2 if and only if d^tf0. Proof. The 3-equicontinuous sets in cpd are the bounded finite-dimensional subsets of q>d. Thus Xlf is the finest locally convex topology on <pd, and 3/ is the topology whose open neighbourhoods ofo consist of those sets whose intersections with the finite-dimensional bounded sets M containing o are open neighbourhoods ofo in M.
9. Grothendieck's construction of the completion 269 The open ^-neighbourhoods of o can aiso be defined as the sets U whose intersections with the finite-dimensional linear subspaces H of cpd are open neighbourhoods ofo in H. Suppose that d = K0. Once again we denote by cpn the linear subspace of cp consisting of the vectors x={£l,...,£„,0,0,...}. Every 3/-neigbourhood of o contains one of the form U=\J U„, where U„ is an open neighbourhood ofo «=i in cpn and U„aUn+1. Ul contains a compact absolutely convex neighbourhood Kx ofo. Since KlczUl, there is a compact absolutely convex neighbourhood K2 ofo with K j + K2 <= U2. Repeating this process we obtain by § 16,1. (3) an absolutely convex ^-neighbourhood ofo Kx + K2+ •••<= U. Xf and Xlf are therefore the same on cp. Now suppose that d>K0. In order to show that the topologies %f and %lf are different on cpd, it is sufficient to give a convex ^/-closed set C which does not contain o, but for which o is a 2^-closure point. We have constructed such a set in § 17,5.(4). We need only set the space E considered there equal to cpd. The set C constructed there intersects every finite- dimensional linear subspace in a closed convex polyhedron, and so it is ^/-closed. Further since the intersection CnC with every convex a-body C containing o is non-empty, o is a 3^-closure point of C. Finally, using an example of Collins [1], we show that in general Xlf is different from X°. Let £ be a (B)-space, and let E be its dual. We consider the dual pair <£,£'> and the topology Z = XS(E) on E. Then by 7.(3) X° is equal to Zb(E% and so is the norm topology on E. The topology Xlf is however the finest locally convex topology on E, and so in general it is different from £° (cf. § 18, 5.(5)). We shall determine the topology %lf more precisely in the next number. 9. Grothendiec k's construction of the completion. So far we have become acquainted with two methods of constructing the completion E[%~\ of a locally convex space £[£], one depending upon the construction of the completion of a uniform space (§ 15, 3.) and the other using the embedding of E[Z~\ in a topological product of (B)-spaces (§18,4.). We now consider a third method, due to Grothendieck [1]; this method has already been used in the linear topological case (cf. § 13, 3.). (1) Suppose that <E2,E1) is a dual pair and that X is a locally convex topology on Ex lying between (XS{E2) and <Xh(E2). Then the dual space of £2p/], and of E2[%lf~\, consists of all linear functional on E2 whose restrictions to the %-equicontinuous subsets of E2 are %S(EX)- continuous. Proof. Suppose that u is a linear functional on E2 whose restrictions to the £-equicontinuous sets M containing o are weakly continuous. If U£ is the set of all xeE2 with |wx|<e, UEnM is an open ^(i^)- neighbourhood ofo in M. By the proof of 8.(1), UE is therefore an open ^/-neighbourhood, and, since Uc is absolutely convex, it is also a
270 §21. The different topologies on a locally convex space ^-neighbourhood of o. u is therefore 3/- and 3^-continuous at o, and consequently continuous throughout E2. Conversely, the restriction of a Zf- or 3^-continuous linear functional to any M is clearly weakly continuous. We now have the following theorem of Grothendieck: (2) Suppose that <£2>£i) *s a dual pair, and that 9JI is a saturated collection'of bounded subsets of E2 which cover E2. Then the completion of Et [3OT] consists of all the linear functionals on E2 whose restrictions to the sets Me9Jt are weakly continuous. Expressed differently, if we form the topology %f or %lf on E2 corresponding to the topology % = %mon E^then (E2[^/]X and {E2[Zif~\y are equal to £, [3]. Proof. [Pelayo Henriques]. Let Ex be the space of all linear functionals on E2 with weakly continuous restrictions to the sets Me9Jt, and let E1 be the completion of £i[3OT]. E1 c Ef, and Et <= E*> by §20,9.(2) and §18.4.(4). ^ We show that E1(^EX. All the MeWl are equicontinuous subsets of F1=E1[2SW]' and by 4.(5) the topologies ^(EJ and Zs(Ei) coincide on M. But then every zeEi <= Ef is a linear functional on E2 c E\ whose restrictions to all the sets M are weakly continuous, so that z lies in Ev Secondly we show that Et is 3^-dense in Ev This implies that Ex cz El5 and so Ex =EV <E2,E1> is a dual pair, and from the definition of Ex the topologies 3s(Et) and ZS(E^) coincide on the sets MeW. Thus the sets MeW are also ^(EJ-bounded, and i^pl^] is locally convex. It is sufficient so show that every veE\=El[Zm]' that vanishes on Et vanishes on the whole of Ev Suppose that v is an arbitrary element of E'v Then there is an absolutely convex MeW with sup|i;z|^l, i.e. ve(M°)°, where the polar zeM° of M°c=E1 is formed in E[. (M°)° is the ^(EJ-closure of M in E'v There is therefore a ^(E^-Cauchy filter g on M with i; = limg. Since ZS(E\) anc* ^s(^i) coincide on M, g is also a ^(EJ-Cauchy filter on M. As such it has a limit v0 in E[, and v0 is the restriction of v to E1? since if xeEu v0x = lim<$x = vx. If now v0=o, g converges on M to oeM. But this also holds for g considered as a ^(EJ-Cauchy filter on M, and so v=o as well. The following completeness criterion is an immediate consequence of (2): (4) The locally convex space E[X~\ is complete if and only if every linear functional on E' whose restrictions to the equicontinuous sets on E' are weakly continuous is weakly continuous on the whole of E'.
9. Grothendieck's construction of the completion 271 If we apply 3.(4), we obtain from (2): (5) // the complete locally convex space E[%~\ contains a countable total subset, for example if E[%~\ is a separable {¥)-space, then every weakly sequentially continuous linear functional on E is weakly continuous. (4) characterises the completeness of a locally convex space in terms of duality. (4) can be expressed in a somewhat different form which goes back to V. Ptak [1] and H. S.Collins [1]: (6) The locally convex space E[%~\ is complete if and only if every Zf-closed linear subspace H of codimension 1 in E' is also weakly closed, i. e. if and only if the fact that H nM is weakly closed for every weakly closed X-equicontinuous subset M of E implies that H is weakly closed. A hyperplane H in E' is %f-closed if and only if it is Xif-closed. We shall prove the last assertion first. Suppose that the ^/-closed hyperplane H is given by z(u) = 0, ze (£')*, ueE, and suppose that u0$H. By 8.(1) there is a circled ^/-neighbourhood U of o with (w0+ U)nH empty. z(U) is circled and, as it is a subset of P (respectively T), it is therefore absolutely convex, so that z(\~ U) = z{U). Since 0$z(w0+L7) it follows that O$z(w0+P U), and so the ^-neighbourhood w0+ r~ U of u0 does not meet H; H is therefore also 3^-closed. It is therefore sufficient to prove (6) with %lf in place of %f. By § 15, 9.(1), the 3^-closed hyperplanes through o are precisely the null -spaces of the ^X'^-continuous linear functionals on E. By (4) the ^-continuous linear functionals are weakly continuous, i. e. their null- spaces are weakly closed, if and only if E[%~\ is complete. We can now determine the topology Hlf more precisely: (7) Suppose that E[X~\ is a locally convex space, and that £[£] is its completion. The topology %if on E' corresponding to X is the topology %° of uniform convergence on the compact subsets of E. For it follows by 3.(2) from {E'[XlJ"])' = E that %lf is the topology Xw of uniform convergence on some saturated collection 9i of bounded subsets of E which cover E. Since by 6.(2) the topologies £s(£) and £S(E) coincide on the £-equicontinuous subsets M of E', %if is also the finest topology 2OT which coincides with ZS(E) on the sets M. But by 7. (2) a) this is Z°. (8) Suppose that £[£] is a locally convex space. The topologies %° and %if on E are the same if and only if every compact subset of E [%~\ lies in the closure of a precompact subset of E [%~\. A base of ^-neighbourhoods of o in E is obtained by taking all the sets C°, where C is absolutely convex and precompact in £[£]; it is also obtained by taking all the sets (C°°)°, where C°° is the compact cover of C in E[%~\. On the other hand the sets K°, with K compact in
272 §21. The different topologies on a locally convex space £[2], form a base of ^-neighbourhoods of o in E'. The two topologies are therefore equal if and only if each K lies in a C°°. 10. The Banach-Dieudonn£ theorem. Interest in the topologies 3/ and %lf goes back to a classical result of Banach about (B)-spaces. A new proof was given by Dieudonne, using a method which can be applied to metrizable spaces. The Banach-Dieudonne theorem states: {I) If E[%~\ is locally convex and metrizable, %° and %f coincide on E'. Expressed in another way, the topology XC{E) of precompact convergence is the finest topology on E' which coincides with the weak topology on all the equicontinuous subsets of E'. Proof. We must show that every open ^/-neighbourhood V of o is a ^-neighbourhood, and so that there is a precompact subset C in E with C° c V. Let Ul => U2 =>... be a sequence of absolutely convex neighbourhoods of o in E[%~] which form a base of neighbourhoods of o. We set U0 = E. We need the following lemma: (2) For each n>0 there is a finite set F„_l cz JJn_ l with the property n-l that if Cn= [j Fp then the set U°n n C° is contained in V. We prove this using complete induction. Suppose that the sets Fp have been determined for p < n in such a way that [/° nC°^V for n- 1 Cn= U FP> P = o We set Dn=Un+1n(E'~V). Un+l is weakly compact, and %f induces the weak topology on U°n+l. Since V is open, E'~V is %f- closed, so that Dn is weakly closed, and is therefore weakly compact. We suppose that there is no finite subset Fn of Un with the required properties. Then for every finite subset F of Un (C„uF)°n Ucn+1 does not lie in V, so that C„ nF° nDn is not empty. All the sets of the form C°nF°n Dn are weakly compact, and they form a filter base on D„, since the intersection of finitely many sets of this form is again of this form. Since Dn is weakly compact all these sets C°nnF° nDn must have a common element w0; this must lie in C°nnU°nc\Dn, which contradicts the relation Ccn n U°n c V. Thus there exists an Fn c U„ with the required property, and so the lemma is proved. 00 We now prove(l). We form the set C= \J Fp. It is relatively compact p=i in E[%], since every subsequence in C converges to o, because of the
11. Real and complex locally convex spaces 273 conditions on the sets Un. Further C° c C°, so that C°nU°n^V for all n. Since E is the union of the sets l/°, C° c K (3) £z;ery precompact subset of a metrizable locally convex space E[X~\ lies in the closed absolutely convex cover of a sequence convergent too in £[£]. The topology %C{E) therefore coincides with the topology of uniform convergence on the sequences which converge to o in E[%~\. This is contained in the proof of (2): %C{E) is coarser than %f, and so, by this proof, given a ^-neighbourhood K° of o, with K precompact in £, there is a set C, consisting of a sequence convergent to o, with C° c K°; by the theorem of bipolars, C°° = T(C) => X. Using the definition of 3/ and the ideas contained in the proof of 8.(1), we obtain the following direct consequence of (1): (4) A subset M of the dual E' of a metrizable locally convex space E[%~\ is %c-closed if and only if Mn5 is weakly closed for every weakly closed equicontinuous subset B of E'. The next result is a special case of (4): (5) A convex subset M of the dual E' of an (F)-space E[%~\ is weakly closed if and only if the intersection M nB is weakly closed for every weakly closed bounded subset B of E'. For, by 6.(1), 3C(£) is coarser than 3fc(£), so that it is compatible with the dual pair <£',£>; by §20. 7.(6), M is ^-closed if and only if it is weakly closed, so that (5) follows from (4). In addition, we record the classical result for (B)-spaces: (6) // E is a (B)-space and if H is a linear subspace of E, then H is weakly closed if and only if H c\K is weakly closed, where K is the closed unit ball in E'. In the separable case we have (7) // E[%~\ is a separable (F)-space, a convex subset of E is weakly closed if and only if it is weakly sequentially closed. This follows from (5), for by 3.(4) every bounded subset of E is metrizable in the weak topology. 11. Real and complex locally convex spaces. We have developed the theory of locally convex spaces in the real and the complex case simultaneously. It is occasionally desirable to be able to carry a theorem proved for real locally convex spaces over to the complex case. We shall prove a theorem which in many cases makes this possible. Suppose that E[X~\ is a complex locally convex space. We can also consider it as a real locally convex space (conversely, it is not always 18 Kothe, Topological Vector Spaces I
274 §21. The different topologies on a locally convex space possible to consider a real locally convex space as a complex one, cf. Dieudonne [8]). Let E' be the complex dual, E'r the real dual. By § 16, 3.(1), Er = ^R{E), i.e. E'r consists of the real parts ul = (iRv of the complex linear functional v = u1+iu2 on E. Conversely v is determined by its real part, for by § 16, 3.(2) u2 is defined by (1) u2x = — u^ix). Consequently on the one hand we have the complex dual pair <£',£>, and on the other the real dual pair (E'r,Er}. Topologies can now be defined on E in two ways; the question of how far they coincide is answered as follows for the most important topologies: (2) // E[%~\ is a complex locally convex space, E' its complex dual and E'r its real dual, then the topologies %S(E') and Hs{Er) coincide, as do %b{E') and %h(E'r\ and Xk(E) and Zk(E'r). Proof. Using the definition of the polar given in §20, 8., it follows that the polar in E'r of M c E is equal to 9t(M°), where M° is the polar in E. Therefore by 3.(1) the £-equicontinuous subsets of E'r are the real parts SR(iV) of the £-equicontinuous subsets N of E'. If 9JI is a total saturated collection of bounded subsets of £', the topology Xm(E) on E is therefore equal to the topology IW(aW)(^). We observe that $R($R) is also saturated. In particular if M is a weakly bounded subset of E', $R(M) is clearly a weakly bounded subset of Er\ conversely if N is a weakly bounded subset of E'r the set of functional v = u1 + iu2 constructed as in (1), with uxeN, forms a weakly bounded subset of E. It therefore follows from 28W = 2W(8W) that Zh{E) = Zh{Er). The £s(£')-equicontinuous sets are tne (complex) finite-dimensional bounded subsets of E, and their real parts are the (real) finite-dimensional bounded subsets of Er\ from this it follows that Zs(E) = Zs(E'r). Finally, if K is complex absolutely convex and weakly compact in E, ${(K) is real absolutely convex and weakly compact in Er. Conversely suppose that Cx is real absolutely convex and weakly compact in Er. The same holds for the set C2 of functional u2 which correspond by (1) to the Mj gCj. By § 20, 6.(5) the closed real absolutely convex cover Kx of Cl and C2 is weakly compact. If we now again use (1) to form the set K of all v = ui+iu2 with uleK1, K is real absolutely convex, and is weakly compact by § 15, 6.(8). Since, further, by the construction of K, K1 = (iR{K)=^(K), K always contains iv if it contains v; by § 16,1.(1), \~ (K) is contained in 2K, and so \~{K) is absolutely convex and weakly compact. Thus Cx lies in the real part of an absolutely convex weakly compact subset of E. Thus it follows from the definition of the Mackey topology (cf. 4.) that £fc(£') and %k{Er) are equal.
1. The duals of subspaces and quotient spaces 275 § 22. The determination of various dual spaces and their topologies 1. The duals of subspaces and quotient spaces. We considered the corresponding problem for linearly topologized spaces in § 10, 8. Suppose that H is a linear subspace of the locally convex space E[X~\. For the sake of clarity, in this and the next number we shall denote the topology which X induces on H by X (up till now it has been denoted by X). We denote the embedding of H(X) into E[X] by /. / is a topological monomorphism of H[X] into E[X~\. As in §10,8., we also introduce the natural homomorphism of E onto H\ this is the mapping N which sends each ueE to its restriction u(0) in H'. It follows from the relation u{0)y = (Nu)y = u(Iy), for all yeH and all ueE, that N is the adjoint of I. (1) Let H[X] be a linear subspace of the locally convex space E[X]. a) The natural homomorphism N of E' onto H' defines an algebraic isomorphism N of E'/H1 onto H'. In this sense, H' can be identified with E'/H1. b) The X-equicontinuous subsets ofH' are the N-images of the X-equi- continuous subsets ofE, or, considered as subsets of E'/H1, the K-images of the X-equicontinuous subsets of E\ where K is the canonical homomorphism ofE onto E'/H1. Proof, a) is proved as in § 10, 8.(1), and N is the mapping determined by the decomposition N = N K. If M is a X-equicontinuous subset of £', so that M <= U° for some absolutely convex neighbourhood U ofo in £, then clearly \uy\= \u{0)y\^ 1 for ueM and yeUnH, so that N(M) <= (U nH)°; the image N(M) is therefore X-equicontinuous in H'. If conversely M' is 4-equicontinuous in H\ then sup |w(0));|:gl yeUnH for all ui0)eM\ where U is some suitable absolutely convex neighbourhood of o in E. By the Hahn-Banach theorem each such u(0) has an extension u to the whole of £, with sup|wx| ^ 1. The set M of all these xeU extensions u of the u{0)eM' is a £-equicontinuous subset of E for which N(M) = M'. Now suppose that H is a closed linear subspace of £[£]. We give the quotient space E/H the quotient space topology corresponding to X; in this and the next number we shall denote this topology by 2. If u is a continuous linear functional on {E/H)\X\ then the equation ux = u x, xeE/H, xex, defines a continuous linear functional u on E[X~\, for it follows from \u'x\<s for xeU that \ux\<t: for xeK(-1)(L/) = U, where K is the canonical homomorphism of E onto E/H. IK*
276 § 22. The determination of various dual spaces and their topologies We call the mapping u = Iu' which we have just defined the natural embedding/of (E/H)' into £'. It follows from the relation u x = u (K x) = (I u) x for all xeE and all u'e(E/H)' that / is the adjoint of K. We have (cf. §10,8.(4)): (2) Suppose that H is a closed linear subspace of the locally convex space E[X] and that the quotient space E/H is given the quotient topology %. a) The natural embedding I of (E/H)' into E' is an algebraic isomorphism of (E/H)' onto H1. In this sense, (E/H)' can be identified with H1. b) Under this isomorphism the X-equicontinuous subsets of (E/H)' correspond to the H-equicontinuous subsets of E' which lie in H1; i.e. the X-equicontinuous and %-equicontinuous subsets of H1 are the same. P r o o f. a) If u' e (E/H)', then, by the definition of /, the linear functional u = Iu vanishes on H, so that u lies in H1. Conversely if ueH1 the equation ux = u'x defines a linear functional u on E/H; this u is continuous on E/H, since it follows from \ux\ <r, for xeU that \u'x\<s for xeK(U). b) If M' is a X-equicontinuous subset of (E/H)', \u'(x)\^l for u'eM' and xeK(U), where U is a suitable open absolutely convex ^-neighbourhood of o in E. Then if u = Iu'eI(M'), |wx|^l for all xeU + H. Thus I{M')^U°nHL. Conversely if M <= (E/H)' satisfies I(M) <= U°nH1 = (U + H)°, then M is contained in (K(U))°. 2. The topologies of subspaces, quotient spaces and their duals. Again let H be a subspace of the locally convex space £[£], equipped with the subspace topology %. We consider a further topology Xm on £, where SR is a saturated class of bounded subsets of E\ and ask which class of bounded subsets of H' = E'/HL produces the induced topology Xm on//. We denote by $R the collection of all sets M = K{M\ where K is the canonical mapping of E' onto E'/H1. (1) Suppose that H is a linear subspace of the locally convex space E[X~\. IfWl is a saturated class of bounded subsets of E\ the induced topology Xm(E') on H is the same as the topology (%$l(E'/H1). In particular 4S(£') is equal to X^E'/H1) on H. Proof, a) If Me TO is absolutely convex, M defines the ^-neighbourhood M° c\H of o, which consists of all yeH with |wy|^l for ueM. But since uy = uy, where u is the coset of u in E'/H1, M° c\H is
2. The topologies of subspaces, quotient spaces and their duals 277 also equal to (K(M))°. Conversely every ^-neighbourhood (K(M))° of o in H, with absolutely convex M, can be written as M° c\H. b) As M runs through the bounded finite dimensional subsets of £', M runs through the bounded finite dimensional subsets of E'/H1, so that XS(E') is equal to X^E'/H1). This can also be seen directly, as in § 10, 8.(2). If the topology Xm in (1) is coarser than the Mackey topology Xk(E'), so that E[Xm~\' = E\ then we can also prove (1) using l.(l)b), by putting X = X<m there. We then obtain that every 3^-equicontinuous subset in E'/H1 is the K-image of a 3^-equicontinuous set, so that in this case 9JI is saturated. If X^ is finer than Xk, then we can only deduce from the proof of (1) that the saturated cover of M is obtained by taking the weak closures in E'/H1 of the sets K(M\ MeWl. Xb(E' /H1) is equal to Xb(E') if and only if every bounded subset of E'/H1 is contained in the closure of a set K(B), with B bounded in E'. Thus Xb(E'/HL) can be strictly finer than Xb(E') (for examples cf. §27.2 and §31, 5.). In the same way Xk(E'/HL) is equal to Xk(E') on H if and only if every absolutely convex weakly compact subset of E'/H1 is the K- image of an absolutely convex weakly compact subset of E'. We shall give a counterexample in 4. In general, therefore, we can only say: The topology X^E'/H1) (respectively X^E'/H1)) on #<=£[£] is finer than Xh(E') (respectively Xk(E% If we apply 1.(2) and interchange E and £', we obtain the following result on the topologies on the dual of a quotient space: (2) Suppose that H is a closed linear subspace of the locally convex space E[X]. Ifffll is a saturated collection of bounded subsets of E, the induced topology X^E) on (E/H)' = H1 coincides with the topology Z&(E/H). In particular XS(E) is equal to XS(E/H) on (E/H), = H1. The remarks following (1) apply in this case, as well. For the topologies on a quotient space we have: (3) Suppose that H is a closed linear subspace of the locally convex space E \X\. Suppose that SR is a saturated collection of subsets of absolutely convex weakly compact subsets ofE', and that 5ft is the collection of subsets o/SR lying in H1 <= E'. Then the quotient topology Xm(E') on E/H coincides with the topology X^H1). $R is also saturated. In particular XS(E') and X^H1) coincide on E/H, and so do Xk(E') and Xk(HL).
278 § 22. The determination of various dual spaces and their topologies Proof, a) By the Mackey-Arens theorem, the dual of £[3^] is also £', and we can therefore apply Theorem l.(2)b) with X = Xyn; the ^-equicontinuous sets in H1 are therefore the ^-equicontinuous sets; i.e. 501 is saturated and X^H1) coincides with *im(E') on E/H. b) The finite-dimensional bounded subsets of H1 are the 4S(£')- equicontinuous sets in H1, and they clearly define the topology ^(H1) on E/H; thus ZS(E') and Z^H1) coincide on E/H. c) By (2), the topologies %S(E/H) and £s(£) coincide on H1, so that the absolutely convex £s(£)-compact sets which lie in H1 are the same as the absolutely convex £s(£/H)-compact subsets of HL\ thus %k(E') and %k(HL) coincide on E/H, by a). Combining (3) and 1.(1) a) we obtain the following result for the topologies of the dual of a subspace: (4) Suppose that H[X] is a linear subspace of the locally convex space E[%]. Suppose that SR is a saturated collection of subsets of absolutely convex weakly compact subsets of E, and that SCR is the collection of all the sets MeSR which lie in H = H11. Then the quotient topology XyjiiE) on H' = E'/H1 is equal to the topology X^(H). ®? is also saturated. In particular, is(E) and ^ts(H) coincide on //', and so do ik(E) and Xk(H). Theorems (3) and (4) need no longer be true for topologies finer than the Mackey topology, and in particular they need not be true for the strong topology. The reason for this is the following: an open absolutely convex £fe(£')-neighbourhood of o in E/H is of the form K(U), where U is an open ^-neighbourhood of o in E. Then K(U)° = (U + H)° U°nHMsa bounded subset B of H'. The polar B° of this in E/H is then the Is(H1)-closure K(U) of K(U). It follows from this that Z^H1) is always coarser than %b(E')onE/H. The ^Xs(H1)-closure of K(U\ however, need not be the same as the S^(F)-closure of K(U\ since %(Ef) need not be compatible with the dual pair {H1,E/H}. Cf. §31,7. for a counterexample. We now give two theorems on the topology of precompact convergence. (5) Suppose that the locally convex space E[X] is quasicomplete, and that H is a linear subspace of E. The topologies £c(£) and HC{H) coincide on H' = E'/H1. Proof. £c(£) is coarser than %k(E), by §21, 6.(1), so that theorem (4) can be applied. But the X-compact subsets of H are the sets which are compact in the induced topology % so that ZC(E) is equal to £C(H). (6) If E [X~\ is an (F)-space and if H is a closed linear subspace of £, the topologies 2,.(£) and %C(E/H) coincide on (E/H)' = H1.
3. Subspaces and quotient spaces of normed spaces 279 For E/H is an (F)-space under the quotient topology % (cf. § 18, 3.(4)), and by the Banach-Dieudonne theorem both ZC(E) and %C(E/H) are the finest topology which defines the weak topology on the weakly bounded subsets of H1, and the weak topologies %(E) and XS(E/H) are the same. (6) can also be expressed as follows: (7) If E is an (F)space, every compact subset of E/H is the canonical image of a compact subset of E. This follows from (6), (2) and one of the remarks following (1), according to which (I is saturated, where (£ is the class of relatively compact subsets of £; (£ therefore contains all the compact subsets of E/H. (7) can also be proved directly using §21,10.(3). If Ux => U2 => ••• is a fundamental sequence of open neighbourhoods of o in E and if xk is a sequence convergent to 6 in E/H, then there is an increasing sequence of integers kn such that xkeK(Un) for /c^/c„, n=l,2,.... There is then a sequence xk in E, with xfcexfc, such that xkeUn for fc^fcn, n= 1, 2,.... The sequence xk defined in this way is therefore convergent to o in E. 3. Subspaces and quotient spaces of normed spaces. In normed spaces the circumstances are particulary simple; indeed in this case the topological isomorphisms are norm isomorphisms. (1) a) If H is a linear subspace of the normed space £, the embedding IofH into E is a norm isomorphism of H into £, and if yeH (2) \\y\\H= sup |M<°>y|= sup M = IM|£. ||u(0)||^1>u(0)6/r ||«||^1,mgE' In particular the topologies £k(F) and Xk(H') on H are the same. b) If E/H is a quotient of a normed space £, the natural embedding I of {E/H)' into E' defines a norm isomorphism of (E/H)' onto H1, i.e. if u e(E/H)' and Iu' =ueH1, (3) I|m'II= suP |m'x|= sup |wx| = ||w||. ||*|| ^l.JceE/ff ||jc|| ^l.xeE In particular the topologies Xb(E) and %b(E/H) on (E/H)' = H1 are the same. Proof, a) / defines a norm isomorphism, by definition. Considered as an element of£ an element yeH has norm sup \uy\, by §17,6.(4), ||«||^l,«e£' and considered as an element of H it has norm sup |w(0)j>|, and these must be the same. "" )|l = 1'"( )eH b) In (3) the suprema can be taken over all ||x||<l and all ||x||<l respectively. If ||x||<l, then ||Xx|| = ||x||< 1; conversely for each x with
280 § 22. The determination of various dual spaces and their topologies ||x||<l there is an x with Kx = x and ||x||<l. It then follows from the relation u'x = u'(Kx) = (Iu')x = ux proved in 1. that the two suprema in (3) are equal. (4) a) IfH is a closed linear subspace of the normed space E, then the norm on E/H satisfies (5) ||jc|| = inf||x||= sup \u'x\. xex Hm'II ^l,u'e(E/f/)' In particular the topologies Zk(E') and %k((E/H)') on E/H are the same. b) // H is a linear subspace of the normed space E, the natural homo- morphism N = N K of E' onto H' defines a norm isomorphism N from E'/H1 onto H'; i.e. if ueE'/H1 and u{0) = NueH' we have (6) ||ii|| = inf|M|= sup |u(0)y| = ||u(0)||. ueu \\y\\£l,yeH In particular the topologies Xb(E) and %b(H) on H' are the same. a) follows from § 17, 6.(4) and § 14,4.(1). We shall prove b). Since u is the coset of all those ueE' whose restriction to H is equal to u{0\ we have ||w|| = sup \u(0)y\ for all ueu, so that ||w|| = inf||w|| = ||w(0)||. Ilyll^i.yetf On the other hand, by the Hahn-Banach theorem, for each u{0) with bound ||w(0)|| on H there is an extension u of equal bound on £, so that we also have ||m(0)||^||m||. 4. The quotient spaces of/1. In Hilbert space theory it is shown that every quotient space of I2 is norm isomorphic either to I2 or to K". Z1 exhibits a completely different behaviour: (1) Every separable (B)-space E is topologically isomorphic to a suitable quotient space of ll. Proof. Let xx,x2,... be a dense sequence of elements in the unit ball of E. We define a mapping A from Z1 into E by making each vec- 00 tor x = (£i)el1 correspond to the element Ax= £ £txt in E. Since i=l oo £ |^| < oo, the series defining Ax in E is convergent and ||>4x|| g £ |^| = ||x||, so that A is a norm-continuous linear mapping. l = 1 Its image space is the whole of E\ given x with ||x|| = 1, we determine 1 a sequence xni, x„2,... 1 1 for which ||x-xni|| <-, 1 °° 1 <-3,...; then £ -kx„k = 1 z fc=i z x-x„, - X. 1 _ 2 X"2 It follows from the Banach-Schauder theorem (§ 15,12.(2)) that A is a topological homomorphism. Consequently E is topologically isomorphic to l1/N\_A~\, where N\_A~\ is the kernel of A.
4. The quotient spaces of Z1 281 Conversely by §14,7.(7) every quotient space ll/H, where H is a closed linear subspace, is a separable (B)-space. (2) Every weak Cauchy sequence in ll is a strong Cauchy sequence, and ll is weakly sequentially complete. The weak convergence is with respect to (/1), = /°°. In order to prove (2) we use the method of the "sliding hump" (cf. §20,11.). First we consider a sequence x(n)ell which is weakly convergent to o. If this were not strongly convergent to o there would be an e>0 and infinitely 00 many np j=\,2,..., for which ||x("-,>|| = £ \^\f,j)\>s. Let JVX be chosen large enough so that £ \£\ni)\^e and X|#ni)|>fe. Numbers JVi + l 1 Ni Ni vl9...,vNl ofmodulus 1 can then be chosen so that ^i;.^ni) = ^|^ni)|>f e; i i then, however the vk are chosen with modulus 1, for k>Nu we have t»it\ni)\>fr>i*. 1 I Ni For the next step we choose nj2 large enough, so that ^|^"^}|^^e. 00 1 Then we determine an N2>Nl so that £ |^"^)|^^£ and iV2 N2 + l Xl£l"j2)l>f£- We can then choose vNl + 1,...,vNl ofmodulus 1 so that i iV2 iV2 X vi^TJl)= X l£i""'2)l>fe; then, however the subsequent vk are JVi + l Ni + l chosen with modulus 1, we have I l^1! I °° I Repeating the process, we obtain a vector u = (vn)elco with |ux("Jk)|>- for all /c=l,2,...; this contradicts the weak convergence of x{n) to o. For a weak Cauchy sequence, we proceed in an analogous way: if it were not a Cauchy sequence, there would be a sequence of pairs of indices {npm^ and an e>0 for which ||i("-/) — x{mj)\\>e9nj9mj->oo. Then, in the proof above, we use the sequence x{nj) — x{mj\ which is weakly convergent to o, instead of x{nj). Since I1 is a (B)-space, it follows that Z1 is also weakly sequentially complete. (3) In I1 the weakly compact, weakly sequentially compact and strongly compact sets coincide. 00 Ivifl"'* 1 All iV2 Z Ni + 1 - Ni z 1 - 00 z iV2+l
282 § 22. The determination of various dual spaces and their topologies A bounded subset C of I1 is strongly relatively compact if and only if 00 lim sup £161 = 0. In particular the closed unit ball of I1 is not weakly compact. Proof. Suppose that M is a weakly compact subset of/1. Since Z1 is separable, /°° is weakly sequentially separable, by §21,3.(5). There is therefore a countable weakly dense subset N in /°°. The topology £s(iV) on I1 is Hausdorff, and it therefore coincides on M with the topology £s(/°°). Consequently M is metrizable under £s(/°°). Thus by §4, 5.(4) every weakly compact subset M of I1 is weakly sequentially compact. By (2), every weakly sequentially compact subset is strongly sequentially compact, and so it is strongly compact, by § 4, 5. (4). But every strongly compact set is certainly weakly compact. This proves the first part of (3). 00 Now suppose that C is the subset of all * = (£,) in ll with £ |£t| f^dn i = n and suppose that dn-+0. If x{k) is a sequence in C, we can obtain a subsequence which converges in each coordinate, by using a diagonal procedure. Suppose then that x{k) is coordinatewise convergent to x(0). 00 00 Then £ l^l^d,, so that £ IS{0)I^4.> i-e- *<0)eC- Further it fol- i = n i=n oo lows from £ \^ik) — ^[0)\^2dn that x(k) converges weakly to x(0). Every i = n such set C is therefore weakly sequentially compact. 00 On the other hand if C is a bounded subset of I1 with sup Y |^| = d„-f»0, C contains a sequence x{"k) with £ |£["k)|^m>0. It is easy i = nk 00 m to obtain a subsequence x{nj) of this with £ |#"-/) —$n/)| ^-r- for all 7,/; C is therefore not strongly compact. We observe that the assertion about the unit ball of I1 has a simpler proof, which will be given in § 23, 5. We now give the counterexample which was announced in 2. By (1) there is a topological homomorphism A of I1 onto I2. By § 20, 9.(5), the unit ball K2 in I2 is weakly compact but not strongly compact. But, by (3), the image of every weakly compact subset M of Z1 is strongly compact, so that there is no weakly compact set which is mapped onto the weakly compact subset K2 of I2. The canonical homomorphism of I1 onto ll/N[A] therefore has the property that there are weakly compact sets in the image which are not the image of weakly compact sets. We now give /°° the topology Zk(lx). Then I1 is the dual of/00. We take H = N[A]L as linear subspace of /°°. By the remarks in 2. preceding (2), the topology ^tk(H') = %k(ll/N[A~\) is strictly finer than the topology Zk(ll) on H. Since ll/N[A] is
5. The duality of topological products and locally convex direct sums 283 topologically isomorphic to the Hilbert space /2, the topology Xk(ll/N\_A~\) is the norm topology on //, which makes H norm isomorphic to l\ It follows from 3.(l)b) that this topology on H is also equal to the topology ^(Z1), and so it is equal to the topology induced on H by the norm topology on /°°. The last statement is true in general: (4) The strong dual of a separable (B)~space is topologically isomorphic to a weakly closed linear subspace of /°°. This follows directly from (1) and 3.(l)b). In § 14, 8. the spaces lld and 1% were defined for arbitrary cardinal d. If £ is a general (B)-space, and if M = {xa} is a set with cardinal d which is dense in the unit ball of E, then E is topologically isomorphic to a quotient space of /j, giving a generalization of (1), and E' is topologically isomorphic to a weakly closed linear subspace of /^. (2) is also true for /j, and the proof is the same. We observe further that every (B)-space can be embedded in a norm- preserving way in a suitable /*, and in particular every separable (B)-space can be embedded in /°°. For this it is sufficient to pick out a weakly dense subset N={u(X} of the unit ball in E\ and to make each xeE correspond to the vector x = (£J9 with £a = wax, in the corresponding l^. In the separable case we can manage with countably many ua, by §21,3.(5). 5. The duality of topological products and locally convex direct sums. Suppose that a collection of dual pairs <Fa,£a>, aeA, is given. The direct sum F = © Fa and the product E = TT £a form a dual pair a a <F,£> in a natural way, when we define the bilinear form ux for two elements u = (uJeF and x = (xa)eE by setting ux = ^waxa, where uaxa is the bilinear form on <Fa,£a>. a The following rules holds for forming polars in <F,£>: (1) // each Ma is a closed absolutely convex subset of Ea and if M°a is its polar in Fa, then the polar (TT Ma)° in F of the subset TT Ma of E is equal to \~ M°. a // each Na is an absolutely convex subset of Fa and if JV° is its polar in £a, then the polar (\~ Na\° in E of the subset f~ Na °f F is equal to a Proof. The first statement follows from §20,8.(10), for TTMa can a be considered as the intersection of the sets M{P) = TT M(/}, where M{P = Mfii and M[p) = Ea for a=#£, and the polar (MiP))° in F is equal to the polar M°p in Fp. The second assertion follows in a similar way from §20, 8.(9).
284 § 22. The determination of various dual spaces and their topologies (2) The dual of a topological product TT£a[jXa] is algebraically iso- a morphic to the direct sum © E'a of the duals; the dual of a locally convex a direct sum © £a[£a] is algebraically isomorphic to the product TT£'a of the duals. a In the dual pairs ( © £a, TT Ea\ and ( TT E'a9 © £a) w/iic/i arise in this way, the bilinear form is given by MX = ^uaxa,uaG£'a,xa6£a. Proof, a) If we(T[£a[£a]Y, w is bounded on a neighbourhood py of o. This can be taken in the form TT Ua9 where Ua. + Ea. for finitely a. many a and Up = Ep for the remaining indices, w clearly vanishes on the subspace TT£«. If P is the projection of TT£a onto TT£a. and Q P a. i the projection onto TT EB9 where each projection maps the other space P to zero, then the relation u x = u(P x) + u(Q x) = u(P x) holds for each x = (xa)eT\Ea. Thus u can be considered as a linear functional on n <* TT £a.[IaJ. The restriction ua. of u to £a. is an element of (£a.|jXa.])', i = 1 since the topology induced on Ea. by the product topology of TT £apXa] is equal to Xar Since TT JSai[IaJ= © £«,[£«,], we have e = 1 i = 1 n n WX = W(PX) = W^ Xa,= X "«^a,» Ma,-e£ar i=l i = 1 Thus each w is mapped in a one-one way to an element of © £a; con- a versely if (wa)e©£a the formula (ua)x = YJuaxa clearly defines a con- tinuous linear functional on TT£a[Ia]. a b) By §18,5.(2), the topology % of the locally convex direct sum © £a[£a] induces the topology Xa on each space Ea9 so that the restric- tion ua of a continuous linear functional uei® £a[£a]J to £a is an element of Fa; clearly if x = (xa)e©£a we have wx = £waxa (the sum a a containing only finitely many non-zero xa). In this way each we(©£a|JXa]Y is mapped in a one-one way to a (ua)eT\E'a. Con- versely if v = {va) is an arbitrary element of TT£^, and if |t;axa|^l a for each xa in the ^-neighbourhood Ua of o in £a, then |i;x| ^ 1 for each x in the neighbourhood \~ Ua of o in © £a[£a], so that v is continuous. a ay Consequently TT£^ is algebraically isomorphic to ©£,[0 . a \ a y We now determine various topologies on topological products and locally convex sums.
5. The duality of topological products and locally convex direct sums 285 (3) Suppose that £pX] = TT£a[£a], with dual space E = @E0L. a a Suppose that a saturated collection SRa of Ea-bounded subsets is given in each Fa. We denote by 9JI the saturated collection of bounded subsets ofE' m consisting of all the finite direct sums © Ma.,Ma.e9Jla., together with i= 1 their subsets. The space E\%^\ is then the topological product of the spaces £a[2anJ- In particular £[IS(F)] = TT£.[!,(£'«)], E\_Xb{E'j\ = TTEa[Xb(E'j]9 E[I^(F)] = n£a[I^(E'a)] and E[Ik(F)] = n£a[Ik(£^]. " a a Proof, a) The fact that %S(E) is equal to the product topology on TT£a[£s(£a)] can be deduced directly from (2) and the definition of the product topology. b) We can take as base of neighbourhoods of o for the topological product TTE^IgnJ the sets TTl/a, where l/ai = M°., Mafe9Ma. for a a finitely many at and Ua = Ea otherwise. By a), TT Ua is weakly closed, since the sets Ua are, so that TT Ua= (n lO°°. By (1) we have (n C/a = P M°°. But the saturated collection of subsets of E defined by the i= 1 ' sets I- M°° is equal to the saturated collection defined by the sets © Ma. i=l * i= 1 n Since the sets © Mai are absolutely convex and weakly closed when i= 1 the Ma. are, the collection of subsets of E which are equicontinuous for the product topology is precisely 2R. c) By § 18, 5.(4) an E-bounded subset of E is always contained in a set n n © B'ai, where Ba. is £a.-bounded in Fa.. Conversely every set © B'a. is weakly bounded in E. The product topology on TT£a[^Xb(Fa)] is therefore equal to Xb(E% by b). d) The topologie X^ which was introduced in §21,5. is the topology of uniform convergence on the strongly bounded sets. Since every strongly bounded subset M' of E is weakly bounded, there are only finitely many non-zero projections M'a = PaM' in the spaces Ea. Every set of the form B = TTBa, where Ba is bounded in £apXa], is bounded a in TT£apa]. It follows directly from sup |wx|<oo that a ueM',xeB sup |uaxJ<oo, i.e. M'a is strongly bounded in Ea. As a result, n every strongly bounded subset of E is contained in a finite sum © M' i= 1 where M'a. is strongly bounded in E'a.. Since conversely every such set
286 § 22. The determination of various dual spaces and their topologies is strongly bounded in E', it follows that 3ft*(E') is equal to the product topology of T\ Ea[Xb*(E'j]. a. e) 3S(£J coincides with the topology induced on E'a by %S(E), so that every absolutely convex 3s(£a)-compact subset Ca cz E'a is ^(incompact, and conversely. If C is an absolutely convex weakly compact subset of E, its projections Ca in the spaces E'a are also weakly compact, and only finitely many of these sets Ca are different from o, since C is n weakly bounded. It therefore follows from this that C cz © Ca, where i= 1 Ca. is weakly compact in E'ai; conversely if the sets Ca. are absolutely n convex and weakly compact, then the set © Ca is again absolutely i= 1 convex (§ 16,1.(3)), and it is weakly compact, by Tychonoff's theorem. It follows from this, by a), that %k(E') is the product topology on T\Ex[Zk(E'Jl a. The corresponding results on locally convex direct sums run as follows: (4) Suppose that E[X] is the locally convex direct sum ©£a[3J, a with dual space E'= T\ E'a. Suppose that a saturated collection 9Jla of a Ea-bounded subsets is given in each E'^ Let 9JI be the saturated collection of subsets of E' consisting of all sets of the form TT Ma, M^yR^, together a with their subsets. Then £[3^] is the locally convex direct sum of the spaces £a[3anJ- In particular we have E[Xb(E')] = © Ea[Xb(E'J], E[Zb*(E')'] = ®Ea [XAE'j] and E [Sk(F)] = © Ea [2fc(E'a)]. a a For the weak topologies, however, the topology of the locally convex direct sum © E^X^E'J] is equal to the weak topology of E if and only a if there are only finitely many summands Ea\%s(E'0f]. Proof, a) The sum topology %' on © Ea [3^J has a base of neigh- bourhoods of o of the form U=[~ M°a = C M°, where each Ma is an a a absolutely convex weakly closed subset in $Ra. By the second part of (1), (rM;M° = TTM°° = TTMa. It follows that Wl is the class of 3'-equicontinuous sets, and H'= %m, provided that we can show that n [/°°-(nMj° = [/. Clearly (TTMa)" => U. Suppose that x=£xa, xa.eEa., is an element of (TT MJ . Then at all events sup |wa.xaJ = pi^l. Since further the u^ can be chosen independently of each
5. The duality of topological products and locally convex direct sums 287 other in such a way that wa.xa.^Pj — s, it follows that ]T pf^l. The element y. = —xa. lies in M°., and consequently x= £ pty, belongs Pi ' i = i to cm:=u. a b) If ©£a[^Xs(E^)] has only finitely many summands, the locally a convex direct sum is equal to the topological product; in this case, therefore, it follows from (3) that the weak topology on E is equal to the locally convex direct sum topology. If there are infinitely many summands and if each 9Jla is the collection of bounded finite-dimensional subsets of Ea, then 9JI contains infinite-dimensional subsets, so that the topology X^ is strictly finer than the weak topology XS(E) on E. We observe however that the formula u x = ]T ua xa (u e E', x e E, ua e Ea, a xae£a) implies that the topology induced on each Ea by XS(E) is the same as XS(EJ. c) If B' is a 3s(£)-bounded subset of E', each of its projections B'a in ££ is a 3s(£a)-bounded subset of ££, so that B' a T\B'a\ conversely every such set TT B'a is weakly bounded in E'. It therefore follows from a a) that the topology Zb(E) coincides with the topology of the locally convex direct sum of the spaces Ea\%h(E'a)]. d) Analogously to (3)d), we can apply § 18, 5.(4) and show that the strongly bounded subsets of E are the subsets of sets of the form TT M'a, a. where each M'a is strongly bounded in Ea. It follows from this that a e) Using the remark in b), we can establish the proof for the Mackey topology in a way which corresponds to (3)e). Theorems (3) and (4) can also be interpreted as statements about the topologies on the duals of locally convex direct sums and topological products. For example it follows from (3) that the weak (respectively strong) dual of ©£a[3j is the topological product of the weak (respectively a strong) duals E^X^EJ] (respectively E'^lX^Ej]). Likewise it follows from (4) that the strong dual of TT Ea [IJ is the a locally convex direct sum of the strong duals Ea[Xb(Ej], and that the dual of TT£a[3J, equipped with the Mackey topology, is equal to the a locally convex direct sum of the spaces E'^X^EJ]. As an example we consider the spaces cod and cpd which were introduced in §15,4. and §18,5.(5) respectively. These spaces are respec-
288 § 22. The determination of various dual spaces and their topologies tively the topological product and the locally convex direct sum of d one-dimensional spaces K. By (2), cod and cpd are dual to each other. By § 18, 5.(6) all the bounded subsets of cpd are finite-dimensional, and by Mackey's theorem (§20,11.(7)) so also are all the weakly bounded sets. Consequently the strong and the weak topologies on cod coincide, and are both equal to the topology of the topological product. The topology on cpd is the topology of the locally convex direct sum, and by § 18,5.(5) it is the finest possible locally convex topology on cpd. Consequently it coincides with the strong topology. cpd and a>d are therefore barrelled (cf. §21,2.). By (3) and (4), all the spaces formed from cpd and cod by repeatedly forming topological products and locally convex direct sums are also barrelled. The duality between topological products and locally convex direct sums established here no longer holds when we replace the latter by the topological direct sums introduced in § 18,5. (5) The dual of the topological direct sum £[X'] of locally convex spaces £a[XJ is equal to the subspace of TT E'a consisting of all u = (ua\ uae E'a with at most countably many non-zero ua. Proof. The topology X' of the topological direct sum is coarser than the topology X of the locally convex direct sum, so that by (2) every continuous linear functional on £[X'] can be represented by a u = (ua)eT\E'a. But such a linear a functional u is only X'-continuous if at most countably many ua are different from o. Let us suppose the contrary. There must be a X'-neighbourhood V= 0 JJa of o, a where each Ua is a ^-neighbourhood of o in Ea, on which u is bounded. For each non-zero ua there is an xaeUa with uaxa = ya>0. It follows that there is a positive 1 k k integer n0 for which yB ^ — for uncountably many /?. But then u Y x*,- ^ — k "o i=l n0 for each element £ xph and these lie in K, for each k. This contradicts the fact [ = i that u is bounded on V. On the other hand, if u = (ua) has at most countably many non-zero components ua, u is X'-continuous on £, by § 18, 5.(8). 6. The duality of locally convex hulls and kernels. We now return to the concepts and results of §19. Suppose that E[Z]=YjA<x(F0C['Z0J) is a the locally convex hull of arbitrary locally convex spaces Fa[Xj. The Aa are then continuous mappings from Fa[Xj into £[X]. If we form the locally convex direct sum ©Fa[Xa] and make each x = £xae©Fa[Xa] a correspond to the element Ax = YJAaxa, A is the topological homo- morphism of ©Fa[Xj onto £[X], by § 19,1.(3), and £[X] is topo-
6. The duality of locally convex hulls and kernels 289 logically isomorphic to the quotient (© Fa[£j)///, where H is the kernel of A. (1) The dual E' of £[£]=£^a(FapXa]) can be represented as the a kernel KA'yl)(F^); in this expression A'a is the adjoint of Aa, mapping a E into Fa'. The mapping A' adjoint to A maps E' isomorphically onto the subspace E of TTFa' consisting of all the elements (A^u). a Proof. A maps ©Fa[£j continuously onto F[£], so that A' maps E into (©FJ = TTFa'. For each x = £xae©Fa and each ueE we have a * (2) M(y4x) = M^y4aXa) = ]T(i4a«) *a = K ") *• a a From this it follows that A'u = (A'au)eT\F^ and that ,4' is one-one. a But by § 19, 6. this means that E = KA^~ ^(FJ, and that 4' is the map- a ^ ping of § 19,6.(1), sending E onto the subspace E of T7F'a. a A completely analogous result holds for locally convex kernels: (3) ThedualEofE[X] = KAi-1)(Fa[Xa]) is equal to the span ]£4a(Fa'), w/zere v4'a 15 f/ze adjoint of Aa, mapping Fa' into E. If A is the topological monomorphism x->(,4ax) ofE\%] into TTFa[jXa], a wi£/z image F, A' is the mapping Yju*^Y*A'aua from ©Fa onto £>4a(Fa). a a a Proof. ,4' clearly maps © Fa' into F'. Since each A'J^FJ is contained a in F', X^a(^a) is at aU events a linear subspace of E. On the other hand a suppose that v is an arbitrary element of E. Since ^ is a topological monomorphism, a continuous linear functional is defined on Fez TTFa[jXa] a by setting u(A x) = vx. By the Hahn-Banach theorem this can be extended to give an element of (TTFaY = © Fa', which we shall again denote by u. Then for this u = Y^u0i and all xeE we have a (4) vx = u(Ax) = Y,K(Aax) = Y,(A'0iuJx = (A'u)x. From this it follows that v = A'u = Y,a'oluv so tnat t;e£/4a(Fa/). Con" a a sequently F' = ]T 4a(i£). The assertion about 4' now follows immediately. a Theorem (1) can also be carried over to inductive limits. 19 Kothe, Topological Vector Spaces I
290 § 22. The determination of various dual spaces and their topologies (4) The dual E of a topological inductive limit E[X]=\imAp<x(Flx[Xlx]) can be represented as the projective limit \imA'pa(Fp), where ApcL is the adjoint of the continuous mapping A pa from Fa[£j into Fp[%p], mapping Fp into F;. Proof. By § 19, 3.(1), if a</?<y we have the relation AyjiAji0L = Aya. This goes over to the relation A'paA'yp = A'ya, and so the relation § 19, 7.(7) is satisfied by the mappings A'Pa from Fp into Fa'. By § 19, 7.(6) we can therefore form lim A'Pa(Fp). By § 19,2, E[2]= (®Fa[Za])/H0, where H0 is the set of all finite linear combinations of elements of the form xa — Apocxa, xaeFa. We observe that Apax0i is an element of Fp. By 1.(2), E can be identified with the subspace Hq of TTFa'. Hq consists of all u = (u0)eT\F^ with a a (5) u(xa-ApaxJ = uaxa-up{Apaxa) = (ua-A'paUp)xa = 0 for all a</J and all xaeFa. This means that ua = A'paUp for all a<j8, and so by §19, 7.(6) H^limyl^'). Theorem (3) cannot be taken over directly to projective limits. We must first take the projective limit in a normal form. By definition E[%] = lim Aap(Fp[Zp]) is the subspace E of TTFa[2j consisting of all (xa) with xa = AapXp for all a</J. We have denoted by Pa(E) the linear subspace of Fa formed by the projections of the elements of £ in Fa. We say that lim A ap{Fp[Zp]) is in reduced form if Pa(E) is dense in Fa[£j, for each a. Every topological projective limit can be taken in reduced form. For this it is only necessary to replace the spaces Fa[£a] by the closures of their subspaces Pa(E), and to replace the mappings AaP by their restrictions to these closed subspaces. The reduced topological projective limit formed in this way is clearly topologically isomorphic to the original one. We now obtain the result corresponding to Theorem (3): (6) The dual E of a reduced topological projective limit E[X] = Mm Aap(Fp[%p]) is the inductive limit \imA'ap(F^), where A'ap is the adjoint of the continuous mapping A ap from Fp[Xp] into Fa[£a], mapping Fa' into Fp. Since the relation A'pyA,lxp = A,ay, follows from the relation AapApy = A%r for a</?<y, the inductive limit limA'ap(F^) = (®F^j/H0 exists, by §19,2. ~*
7. Topologies on locally convex hulls and kernels 291 E[%] is a subspace of TTFa[IJ. By 1.(1), E is equal to (®F^)/E1; our theorem is therefore established if we can show that E1 = H0. n Suppose that u = ]T ua. is an arbitrary element of © Fa', and that x = (xa) is an element in E. If /?^af, f=l,...,n, we have x^^.^Xp for i=l,...,w. Thus we have (7) UX= XM«iX«i = ZM«i(^«i/JX/j) = ( Z^«f/>M«<W i=l \i=l / If ueE1, it therefore follows from (7) and the fact that E is in reduced n form that the equation ]T /4'a{/?Ma.=o holds. By §19,2.(8) this means i = 1 that w belongs to //0. Conversely, it follows in a similar way from (7) and §19, 2.(8) that HqCzE1. 7. Topologies on locally convex hulls and kernels. We first obtain a simple lemma about the formation of polars. (1) Suppose that A is a weakly continuous linear mapping from El [£x] intoE2\^2\' Then if M is any subset of E± (2) A(M)0 = A,{-l)(M°). Proof. A' maps E'2 into E\. By definition A(M)° consists of all the veE2 with Wi;(i4x)^l for all xeM. This is equivalent to 9l(^'i;)x^l for all xeM, and so to A'veM% i.e. veA,{-l\M°\ We now consider the locally convex hull E\%] = YaAa(Fa\%0^\) = a(® ^a[2j). It follows directly from (2) and 5.(1) that (3) //', for each a, Na is an absolutely convex subset of Fa and N° is its polar in F'a, then the polar of A (\~~ Na) in E' is equal to A'{~1) (TT N°\ We can obtain information about the hull topology from this: (4) Suppose that £[!] = Z>la(Fa[IJ) = A (® Fa[xA Let Wla be the collection of %a-equicontinuous subsets of F'a, and let 9JI be the saturated collection of subsets of TT F^ consisting of all the sets TT Ma, a a Mae^Ra, together with their subsets. Then ,4'(_1)(9Jl) is the collection of %-equicontinuous subsets of E\ so that Z = ZA.(- 1)(OT). The neighbourhoods of o of the form A [T~ LfX where each Ua is a ^-neighbourhood of o in Fa[Xa], form a base of ^-neighbourhoods
292 § 22. The determination of various dual spaces and their topologies of o in E[%~\. Thus the sets A (f~ l0° and their subsets form the collection of ^-equicontinuous subsets of E. The assertion now follows from (3). We now determine the equicontinuous sets for the kernel topology: (5) Suppose that E[Z] = KA[-'\Fa[Za]) = A^^(UFa[zA Let ^ be the collection of Ia-equicontinuous subsets of F'a, and let Wl be the n collection of subsets M= © Mat, Ma.e9Wa, of © F'a, together with their i=l ' a „ subsets. We denote by A\W) the collection of the sets A'(M)= £ A'a.(Mai) in E'. i = i Then A'($R) is the collection of X-equicontinuous subsets of E, so inai /-L == ^/I'tyji)' A base of ^-neighbourhoods of o is given by the finite intersections of sets ^_1)(M°), where we can suppose that the Mae^Ra are weakly compact and absolutely convex. The collection of ^-equicontinuous subsets of E is therefore formed by the weakly closed convex covers of finitely many A(~l\M°a)0, together with their subsets. The mapping A'a from F'a into E is weakly continuous, by §20,4.(6); applying (2) to it, we obtain A{-l)(M°a) = A'a(Ma)0; consequently A{-1)(M°a) = A'a(Ma)00, and since Ma is absolutely convex and weakly compact, ,4'a(Ma)°° = A'a(Ma). Since the sets A'^MJ are equicontinuous, so also are the n sets ]T A'ai(Ma), which are absolutely convex and weakly compact, ; = i so that every equicontinuous set lies in such a set. If each $Ra is the collection of finite-dimensional bounded subsets of F'a, A'(W) is the collection of finite-dimensional bounded subsets of E, and so it follows from (5) that (6) The weak topology %S(E) of a kernel K^-^pIJ) is the kernel topology of KAi-l\Fa[X8(F'j]). Interchanging the two spaces, we obtain (7) The weak dual of a locally convex hull E\%\=Y^Aa(Fa\li^\) is equal to KA'<-l)(F£t8(F3]). a There are no corresponding general results for the topologies Xk, Xb and Zb* on a kernel. Suppose that we are given a kernel in the form K^_1)(Fa[^a]), so that we have a subspace of TTFa[IJ. Then a a E = © F'J/H, and A' is the canonical mapping K from © F'a onto E. If y$la is the class of all bounded subsets of F'a, then $R is the class of all bounded subsets of © F'a. The sets M of $R are mapped by K into
7. Topologies on locally convex hulls and kernels 293 bounded subsets of E, but the sets K(M) need not determine all the bounded sets of E (cf. 2.). We can draw similar conclusions about Xk and %b*. In general, therefore, we can only say that the strong topology of a kernel is always finer than the kernel topology of KA[~1\Fa[%h(F,J]). Similar results hold for %k and £b*. There is no result analogous to (6) for locally convex hulls. We saw in 5.(4) that even for locally convex direct sums the weak topology can be strictly coarser than the corresponding sum topology. In general the hull topology of X^a0Fa[£s]) is finer than the weak topology of the hull. However, a (8) Suppose that E[X] = ]£>4a(Fa[Ia]) is a locally convex hull Then we have £[!*(£')] = I^a(Fa[2fc(Fa)]). We can take E[2] in the special form (©Fa[£j)//J, where A = K, the canonical mapping of ©Fa[2a] onto £. Then E is equal to the a space H1 orthogonal to H in TTFa. The mapping K' is the embedding a of E into TTFa. Now let 9Jla be the class of absolutely convex weakly a relatively compact subsets of Fa. The collection 9JI constructed as in (4) contains all such subsets of TTFa, by 5.(4). The collection K'(_1)(9Jl) a consists of all MnH1, with MeWl. H1 is weakly closed in TTFa, and the topology XS(E) coincides with %S(®F^. K,(-1)(9Jl) therefore consists of all the absolutely convex relatively £s(£)-compact subsets of E; the assertion now follows by applying (4) to the compatible topology Z = Xk. The corresponding statement about strong topologies does not hold in general, and indeed it is false in the special case of a quotient space (cf. 2.). The result on the Mackey topology of the dual of a locally convex kernel which corresponds to (8) is true, provided that just one further condition is imposed. We can suppose that the locally convex kernel is a subspace £=K^["1)(Fa[IJ) of TTFa[3:J. The hull topology of a a , £Ka(Fa[2j) is then defined by the neighbourhoods f£nTTCj°, a a where each Ca is a weakly compact absolutely convex subset of Fa[IJ. TTCa is certainly weakly compact and absolutely convex in TTFa, but a a EnT\Ca need not be weakly compact in E, if E is not closed in TTFa. a a If this is the case, however, then Xk(E) is equal to the hull topology of Y,A'(F^[Zk]); in general this hull topology is finer than Zk(E). a
294 § 22. The determination of various dual spaces and their topologies For projective limits, therefore, we have, by § 19,10.(3) and 6.(6): (9) // £[jX]=limAap(Fp[Zp]) is a topological projective limit in reduced form, the topology %k(E) of the dual is equal to the hull topology of \imA'xP(F:[Xk-]). The fact that the inductive limit is a topological inductive limit follows' from the fact that the weakly continuous mappings A'aP are also ^-continuous, by §21, 4.(6).
CHAPTER FIVE Topological and Geometrical Properties of Locally Convex Spaces We continue with the general theory of locally convex spaces. In § 23 the bidual space is introduced, and the question of the semi-reflexivity or reflexivity of a locally convex space is raised. A number of criteria are established, and we examine the connection with other structural properties. In § 24 we consider the question of finding conditions for a subset of a locally convex space to be weakly compact. The important theorems of Eberlein and of Krein are obtained in full generality, as are several related criteria. We give two proofs of Krein's theorem. The first is due to Grothendieck and uses integration theory, while the second was given recently by Ptak, who succeeded in proving the result without the help of integration theory. We also establish three criteria of Klee's for semireflexivity, which are of a geometric nature. § 25 is devoted to a group of problems centred on the Krein-Milman theorem; it deals with extreme points and extreme rays of compact and locally compact convex sets. Closely related to this is the investigation, in § 26, of the various refinements of the concept of convexity. These have proved to be important above all for the structure of normed spaces. Thus the reflexivity of a Banach space follows from the uniform convexity of its unit ball. Strict convexity is of importance for questions of approximation. The concepts dual to strict and uniform convexity, namely smoothness and uniform smoothness, are equivalent to differentiability properties of the norm. § 23. The bidual space. Semi-reflexivity and reflexivity 1. Quasi-completeness. The locally convex spaces met with in applications are frequently not complete. We have already seen that important results in the general theory can be established under the weaker hypotheses of sequential completeness or quasi-completeness (e. g. the Banach-Mackey theorem). Both these concepts were introduced in § 18, 4. We shall meet examples of quasi-complete spaces which are not complete later on. We now given an example of a sequentially complete space which is not quasi- complete. Suppose that d>N0, and that H is the linear subspace of cod consisting of all vectors x = (£g) with only countably many non-zero coordinates £a. H is dense in cod in the topology introduced on cod in § 15,4., and indeed every element
296 § 23. The bidual space. Semi-reflexivity and reflexivity of (Dd is a closure point of a bounded subset of H (cf. § 15,6.); H is therefore not quasi-complete, although it is sequentially complete. Clearly we have (1) Every closed linear subspace of a sequentially complete (respectively quasi-complete) locally convex space is again sequentially complete (respectively quasi-complete). The corresponding assertion for quotient spaces is not true; indeed, the example (pa>(&(jL>(p of § 13,6. shows that the quotient of a complete locally convex space need not be sequentially complete. We observe that the proof given for linear topologies in §13,6. also holds for the locally convex topology on (pco®cQ(p which is obtained from the topologies of co and q> by forming the topological product and locally convex direct sum. (2) The topological product and the locally convex direct sum of sequentially complete (respectively quasi-complete) locally convex spaces are again sequentially complete (respectively quasi-complete). Proof. If £[2] = TT£a[2J, a sequence x{n) = (x(an)) in £[2] is a a Cauchy sequence if and only if the x{"] form a Cauchy sequence in Ea, for each a. The sequential completeness of E therefore follows from the sequential completeness of the spaces Ea. A set B cz £[£] is bounded if and only if each of its projections Ba = PaB is bounded m £a[£J. If the spaces £a[£j are all quasi- complete, the closure Ba of Ba in Ea is complete, for each a, and so also is TT Ba in E. Since this set contains B, B is also complete, provided that it a is assumed to be closed. Applying §18,5.(4), we easily obtain the assertion about locally convex direct sums. A subset M of a locally convex space E[X~\ is said to be quasi- closed if it contains all the closure points in E of its bounded subsets. The intersection of arbitrarily many, and the union of finitely many, quasi-closed subsets are again quasi-closed. If A is a continuous linear mapping from E[X] into F[T], the inverse image A{~l)(M) of every quasi-closed subset M of F is quasi-closed. The quasi-closure M of a set M <=£[£] is the intersection of all the quasi-closed subsets of E which contain M. Each point of M is called a strict closure point of M. Such a point is naturally also a closure point of M, but the converse is not always true. On the other hand a strict closure point of M need not to be closure point of a bounded subset of M (cf. the example at the end of this number). The continuous image A x of a strict closure point x of M is a strict closure point of A(M\ for if A(M) lies in the quasi-closed set N, M
2. The bidual space 297 lies in A{~l)(N\ and ^(_1)(iV) is quasi-closed and therefore contains x, so that A x lies in N. The quasi-completion £ of a locally convex-space E[X] is the quasi-closure of £ in the completion E[X~\. E is quasi-complete under under the topology X induced on E by X. (3) The dual E of a barrelled space E\X\ is weakly quasi-complete. For the bounded weakly closed subsets of E are weakly compact, by § 21,4.(4), and so they are weakly complete. We shall prove the converse in 6.(4). Hilbert space is weakly quasi-complete, by (3), but is not weakly complete (cf. §20,9.(2)). (4) Let Abe a continuous linear mapping from E\_X~] into F\_X'~\. If F is quasi-complete (respectively complete), A can be extended in a unique way to a continuous linear mapping from the quasi-completion (respectively completion) of E[X~\ into F[_X'~\. Proof. Suppose that F is complete. By §15,2.(4), A is uniformly continuous, and so by §5,4.(4) it can be extended continuously to E in a unique way. The extension A is again linear. We have therefore proved the assertion when F[X'~] is complete. If F is quasi-complete, A maps the quasi-completion E into F. But every strict closure point of E in E is mapped by A into a strict closure point of A(E) in F. Since, by hypothesis, this lies in F, A(E) a F. Example. Using the terminology of § 13, 5., the locally convex sum of countably many spaces co is denoted by q>co. The elements of <pco are of the form x = (£ik) 00 = Z £ik cifc» with £ik = 0 for i ^ i0 and all k. Let H be the linear span of the elements i,k=l anfc = ein + cnfc» n,fc=l,2,.... None of the elements cln lies in Hy although cln is the limit of eln + e„fc as /c->oo, and so belongs to the quasi-closure H. Every sum m oo m Yj^in also lies in H; x0= ]T cln is limit of the terms £eln, and so x0 is an element 1 ^ n=l 1 of H. x0, however, is not a closure point of a bounded subset of H; for such a set can only contain linear combinations of the cln + c„fc with nf^n0. x0 is therefore a strict closure point of H, but not a closure point of a bounded subset of H. 2. The bidual space. If we give the dual E of a locally convex space E\X~\ a topology Xw, where 9Ji is a total saturated class of bounded subsets of £[£], then we know from §21,4. that the dual of £'[3^] coincides with E if and only if X^ is coarser than the Mackey topology Xk(E) and finer than the weak topology XS(E). By §21,1.(2), £'[3^] is locally convex, and its dual is a subspace of the algebraic dual (£')*> which, by §20,9.(2), can also be considered as the weak completion E[_XS(E')~\ of E. The dual of Fp-an] can be determined in the following way:
298 § 23. The bidual space. Semi-reflexivity and reflexivity (1) If Wl is a total saturated class of bounded subsets of the locally convex space E[X~\, the dual of £'[3^] is equal [j M, where M is the weak closure of M in (£')* = E[ZS(E')]. Me9W Proof. Every continuous linear functional z on £'[3^] is bounded in modulus by 1 on some suitable 3^-neighbourhood M° of o, where M is absolutely convex and belongs to 501. In other words, z lies in the polar M°° of M° in (£')*. Conversely every element of M°° is a continuous linear functional on E. By the theorem of bipolars, M°° = M, and so we have {E[Z^])' = \J M. Mean (2) // Xm is finer than £s(£), E is a subspace of (£'[3^])'. For each x0eE is a weakly continuous, and a fortiori 3^-continuous, linear functional on E. The finest of the topologies 3^ on E is the strong topology 3^(£), and this therefore produces the largest dual space. In conformity with the terminology for normed spaces introduced in § 14, 5., we denote the dual of E'[Zb{E)] by £", and call it the bidual space of £[3:]. In this case, (1) becomes: (3) // £[3Q is a locally convex space, the bidual E" is the union of the weak closures in (£')* of the boundel subsets of E[Z~]. E" therefore always lies in the £s(£')-quasi-completion of E. Every bounded subset B of E[X~\ is weakly precompact, by § 20, 9.(3). Since (£')* is weakly complete, by §20,9.(2), the weak closure B of B in E" is weakly compact; it therefore follows from (3) that (4) Every bounded subset of E\fX\ is weakly relatively compact in the bidual E". This result can also be expressed as a statement about the topologies on E: (5) The Mackey topology 3^(£") on E is always finer than the strong topology %h{E) on E. 3. Semi-reflexivity. In § 14,5. we said that a (B)-space E was reflexive if the bidual E" coincided with E as a (B)-space. This means first that E" is equal to £ as a vector space and secondly that the norm of E" coincides with the norm of E. Because of § 17, 6.(3), the second statement is a consequence of the first. In the case of an arbitrary locally convex space the corresponding circumstances are more complicated; for this reason we shall begin by considering the general form of the first statement. Suppose that E\%~\ is locally convex. When we consider E as a space of continuous linear functionals on £', we have EczE". £[3T] is said to be semi-reflexive if E is equal to E". We have the following criterion:
3. Semi-reflexivity 299 (1) A locally convex space E[X] is semi-reflexive if and only if every bounded subset of E[%~] is relatively weakly compact, L e. if and only if the topologies Zb(E) and %k{E) coincide on E'. Proof. If E" = E, then the relatively weakly compact subsets are the same as the bounded subsets, by 2.(4), so that %b(E) is equal to %k(E) on E'. On the other hand, if this is the case, it follows from the Mackey- Arens theorem that E" = (E \Zk(E)~\) = E. A second criterion is (2) A locally convex space E[X~\ is semi-reflexive if and only if it is weakly quasi-complete. Proof. Suppose that E is weakly quasi-complete. By §20,9.(3), every bounded set is weakly precompact, and so, since it is contained in a bounded weakly complete set, it is weakly relatively compact. £[£] is therefore semi-reflexive, by (1). Conversely every semi-reflexive space is weakly quasi-complete, by (1), or 2.(4). For (B)-spaces, it follows from (1), (2) and the fact that reflexivity and semi-reflexivity are the same that (3) A (B)-space E is reflexive if and only if the closed unit ball of E is weakly compact, or weakly complete. From (1) and § 21, 2.(2) it follows at once that (4) The strong dual of a semi-reflexive locally convex space is barrelled. The converse of this is not true. For example the (B)-space c0 is not reflexive, by § 14, 7., whereas its strong dual Z1 is barrelled, since it is a (B)-space. (5) Every closed linear subspace H of a semi-reflexive space E[X] is semi-reflexive. The topology £s(//') coincides on H with the topology induced by ZS(E'), by § 22, 2.(1). Since H is weakly closed in E, every weakly closed bounded subset of H is weakly complete, by (2), so that H is weakly quasi-complete, and is therefore semi-reflexive, again by (2). A quotient space of a semi-reflexive space, and likewise the strong dual of a semi-reflexive space, are in general not semi-reflexive (cf. 5. and 6.). (6) The locally convex direct sum and the topological product of semi- rejlexive locally convex spaces are again semi-reflexive. Proof. Suppose that E\_X~\ = @ Ea[2J. By § 18, 5.(4), each bounded <x n subset of E[Z~\ is contained in a set of the form B = © Ba.9 where Bai
300 § 23. The bidual space. Semi-reflexivity and reflexivity is bounded in £ai[3^J- By hypothesis we can suppose that the sets Ba. are £s(£^)-compact. ByTYCHONOFF's theorem, B is compact in the prod- n uct topology of © £a.pXs(£a.)], and this coincides with the topology i= 1 induced by %S(E'), by §22,5.(3). The semi-reflexivity of E[X] now follows from (1). b) The semi-reflexivity of a topological product TT£a[£a] follows a in an analogous way from the fact that every bounded set lies in a product TT £a of bounded and weakly compact sets, and this is weakly compact in TT£a, by §22, 5.(3). a (7) The projective limit E[_X~] = lim Aap(Fp[_Xp]) of semi-reflexive spaces Fp[Xp~\ is semi-reflexive. By § 19,10.(3), E[Z~] is a closed linear subspace of TTFa[£a], and so the assertion follows from (5) and (6). 4. The topologies on the bidual. The bidual E" of a locally convex space E\%~\ forms a dual pair <£",£'> with E\ and so the usual topologies can be defined on E". Since we must consider the original dual pair <£',£> as well as <£",£'> in what follows, we shall occasionally depart from our present usage by mentioning the second space of the dual pair when we describe a topology; for example, %b(E\E") denotes the strong topology on E" with respect to the dual pair <£',£">. Its equicontinuous sets are the £s(£")-bounded subsets of £'. By the strong bidual of E\Z~\ we mean the bidual E" equipped with the topology Zb(E\E")\ the strong bidual is therefore the strong dual of E'[Xb(E)l E is a subspace of E". We now investigate the topology which Xb(E, E") induces on E. This need not be the strong topology on £; rather, we have (1) %b(E\E") induces the topology Zb*(E\E) on E. Proof. By Mackey's theorem (§20,11.(7)), applied to E'[Zb(E)] and its dual £", the ^(E'^E^-bounded subsets of E' are the same as the bounded subsets of E'\Zb(E)\ and these are the strongly bounded subsets of E' with respect to the dual pair <F,E>. Thus the %b(E\E")- and ^IM(^^)-ecluicontinuous subsets of E are the same. Besides the topology Zb(E',E"), there is a second important topology, the natural topology Xn(E'\ on the bidual of £[£]. This is defined as the topology of uniform convergence on the ^-equicontinuous subsets of E'. A base of neighbourhoods ofo for Zn(E') therefore consists of all the polars U°° in E" of the sets U° cz £', as U runs through the absolutely convex neighbourhoods ofo of E[_Z~\.
4. The topologies on the bidual 301 Thus if {U} is a base of ^-neighbourhoods of o in £, the XS{E',E")- closed convex covers in E" of the sets U form a base of ^(^-neighbourhoods of o in E". Clearly, (2) The natural topology Xn(E') always induces the original topology X on E. Xn(E') is always coarser than the strong topology Xb(E',E"), for every £-equicontinuous subset of E' is £fo(£)-bounded, by § 21, 5.(1). When do the natural and the strong topologies coincide on E" ? This happens if and only if the £-equicontinuous subsets of E' are the same as the sets which are strongly bounded with respect to £, and thus if and only if X is equal to Xb*(E'). A locally convex space E[X] is said to be quasi-barrelled if X coincides with Xb*(E). Since Xb*(Ef) is always finer than X and coarser than Xh(E'\ a barrelled space is always quasi-barrelled; the converse is not always true. Later we shall investigate barrelled and quasi-barrelled spaces in detail; for the present we shall content ourselves with a simple characterization of quasi-barrelled spaces. We say that a subset M of a vector space absorbsasetNifpNczM for a suitable p>0. Using this terminology, we have (3) A locally convex space E[X~] is quasi-barrelled if and only if every barrel in E which absorbs all the bounded sets of E[X] is a X-neigh- bourhood of o. For the polar of such a barrel is strongly bounded in E'\ conversely the polar of an absolutely convex strongly bounded subset of E' is a barrel with the given property. It follows from the remarks made above that (4) The natural and strong topologies of a bidual space E" are the same if and only if E[X] is quasi-barrelled, and so if and only if the X-equicontinuous subsets of E are the same as the Xb(E)-bounded sets. By §21, 5.(3), all metrizable locally convex spaces, and in particular all (F)-spaces, are quasi-barrelled. Thus the natural and strong topologies coincide on their biduals. The strong bidual of a normed space is a (B)-space. We shall prove in § 29,2. that the strong bidual of a metrizable locally convex space is an (F)-space. For the present we establish the weaker result: (5) // E[X] is metrizable, the strong bidual E" is also metrizable. For if Un, n=l,2,..., is a base of neighbourhoods of o in E[X~\, the bipolars U°n° in E" form a base of neighbourhoods ofo for the natural, and so for the strong, topology on E".
302 § 23. The bidual space. Semi-reflexivity and reflexivity As a third topology on E" we consider the Mackey topology Zk{E\ E"). It has as equicontinuous sets the absolutely convex £s(£")-relatively compact subsets of £', together with their subsets. Since E" => £, the topology %S(E") is finer than %S(E) on £'. Every £s(£")-compact subset of E' is therefore also £s(£)-compact. Consequently we have (6) The topology %k(E\E") induces a topology on E which is coarser than Xk(E',E). We now show by giving an example that the two topologies can be different. In the strong dual Z1 of c0 the closed unit ball K is Xs(c0)-compact, by §20,9.(5), but it is not Xs(/°°)-compact, by § 22,4.(3). As a result, X^/1,/00) induces a weaker topology on c0 than X^/1,^). 5. Reflexivity. A locally convex space E[X] is said to be reflexive if the bidual E" is equal to £, and if the topology %b{E\E") coincides with the original topology; put another way, E\%~] is reflexive if the strong bidual of E[Z~] coincides with E[X~\. For (B)-spaces, this agrees with the previous definition. A first criterion for reflexivity is given by (1) A locally convex space is reflexive if and only if it is semi-reflexive and quasi-barrelled. This follows directly from 4.(2) and 4.(4). From 3.(2) we obtain (2) A locally convex space is reflexive if and only if it is weakly quasi- complete and quasi-barrelled. The next result contains two further criteria: (3) A locally convex space E[%~] is reflexive a) if and only if X is the Mackey topology and every bounded set in E[X~\ and in E'[Xk(E)~\ is relatively weakly compact, and b) if and only if % is the Mackey topology and E\1L\ and E'[Xk(E\\ are weakly quasi-complete, or semi-reflexive. Proof. Suppose that E[X] is reflexive. From the fact that E" = E, and from the definition of Zb(E',E"), it follows that X coincides with the topology Xh(E'). Consequently the Mackey topology, which lies between X and Xh (£'), coincides with them both. It follows from 3^(E') = Xk(Ef) that every bounded subset of F[£k(£)] is weakly relatively compact, and also that E' is weakly quasi-complete. It follows from 3.(1) and 3.(2) that E also has the same properties. If, conversely, conditions a) or b) are satisfied, then by 3.(1) and 3.(2) E\%\ is semi-reflexive, and further Xb(E') = Xk(E') = X on E. Using the fact that every (F)-space is quasi-barrelled, the following generalization of 3.(3) now follows from (1) and (2):
5. Reflexivity 303 (4) An (F)-space £[£] is reflexive if and only if it is semi-reflexive, and if and only if it is weakly quasi-complete, and if and only if every bounded subset of E[X~\ is weakly relatively compact. The next result follows directly from the definition of reflexivity: (5) // E[%~\ is reflexive, the strong dual E[Xb(EY\ is also reflexive. In many cases, the reflexivity of E can be deduced from the reflexivity ofF: (6) Suppose that £[£] is quasi-complete, and that % is the Mackey topology. If the strong dual E[%b(E)~] is semi-reflexive, E[X~\ is reflexive. Proof. The £6(£)-bounded sets of E' are the £"-bounded sets of £', where E" is the strong bidual. It follows from the semi-reflexivity of E' that these sets are relatively £s(£")-compact. Since XS(E) is coarser than ZS(E") on £', these sets are also relatively £s(£)-compact. But the fact that the ^(E)-bounded sets of E' are the same as the relatively £s(£)-compact sets means that E\%~\ is quasi-barrelled. By (1), our theorem is proved if we can show that E is semi-reflexive. By 2.(3), E" is obtained from E by taking the £s(£')-closure points in (£')* of the bounded subsets of E. Since these can be taken to be absolutely convex, it is sufficient by §20,7.(6) to consider the £k(£',£")-closure points. But, as the argument of the first part of the proof shows, %k(E,E") coincides on E with Xk(E, E) = X. It therefore follows from the assumption that E[X] is quasi-complete that E" = E. In particular, because of (5) we have (7) An (F)-space is reflexive if and only if its strong dual is reflexive. From the fact that c0 is not reflexive (cf. § 14, 7.) it therefore follows that neither Z1 nor /°° is reflexive. By 3.(3) the closed unit balls of Z1 and of /°° are not weakly compact. The strong dual of a (B)-space is again a (B)-space, and so from (7) we obtain (8) // the (B)-space E is not reflexive, all the iterated strong duals are not reflexive, and in the sequences £c£" c£"" <=••■ and E cz E" <z • • • each space is a proper closed subspace of the one following it. In § 29, 2. we shall prove that this holds for (F)-spaces as well. The following analogue of 3.(6) holds: (9) The locally convex direct sum and the topological product of reflexive spaces are again reflexive. Proof. On each of the reflexive spaces £a[£j, Za is the Mackey topology, and so the locally convex direct sum topology X on E[X~\ = 0 £a[2a] is the Mackey topology on E, by § 22, 5.(4). By 3.(6), E[%~]
304 § 23. The bidual space. Semi-reflexivity and reflexivity is semi-reflexive. The spaces E'a[Zb(Ej] are semi-reflexive, and by 3.(6) so also is (E[Z~])' = Tl E'a. The reflexivity of E[X~\ now follows from (3)b). The argument for the topological product is similar. It follows from 3.(5) and (4) that (10) Every closed linear sub space of a reflexive (F)-space is reflexive. For (B)-spaces, we have the further result: (11) If E is a reflexive (B)~space and if H is a closed linear subspace of £, then the (B)-space E/H is also reflexive. For under the canonical mapping K from E onto E/H the image K(B) of the weakly compact unit ball B of E is ^(H^-compact, by § 22, 2.(3). On the other hand K(B) is dense in the closed unit ball of E/H, and so must coincide with it. The assertion now follows from 3.(3). Later (§31, 5.) we shall give an example of a reflexive (F)-space which has a non-reflexive quotient space. (10) is not true for arbitary locally convex spaces, either. We now give an example of this. By (9), the spaces of countable degree, which are obtained from the ground field K by repeatedly forming locally convex direct sums and topological products (cf. § 13, 5.), are all reflexive; in particular the space q>(D®(Dq> and its dual axpt&cpw are reflexive. In §13,6., closed linear subspaces Hx a (paKQaxp and H2<^u>(p ®(poj where constructed with H2=Hl. The quotient space (q>(D®(Dq>)IHl contains a sequence which is certainly a Cauchy sequence for the topology XS(H\\ but which does not converge. It follows from this that (qxo^oxpyHx is not even semi-reflexive. Further H2 is a closed linear subspace of axpQxpco whose dual space coincides with ((pcQ®cQ(p)/Hl9 since H2=Hl. Since this is not semi-reflexive, H2 is not reflexive. In general a non-reflexive (B)-space E has infinite co-dimension in £", as c0 does in /°°, for example. James [1], [2] has given an example of a real (B)-space E with one-dimensional E"/E; further E is norm-isomorphic to E". Civin and Yood [ 1 ] have made a thorough investigation of (B)-spaces with finite dimensional £"/£, which they call quasi-reflexive, relating their properties to ideas in a paper by Diximier [1], which contains interesting results about the iterated duals of a (B)-space. 6. The relationship between semi-reflexivity and reflexivity. The question of whether a locally convex space E\%\ is semi-reflexive depends only on the dual pair <£',£>, since the strong topology Zb(E) on E' is determined by the dual pair alone. Thus % can be replaced by any topology lying between %S(E') and %k{E') without affecting the assertion that E is semi-reflexive. Thus if a dual pair (E2,EX)> is given, Ex can be defined to be semi- reflexive with respect to E2 if (E2[Xh(El)])' = Ex. Ex is then semi- reflexive in the earlier sense with respect to any compatible locally convex topology. In the same way the concept of reflexivity can be defined for dual pairs: a dual pair <£2,£1> is said to be reflexive if Ex is semi-reflexive
6. The relationship between semi-reflexivity and reflexivity 305 with respect to E2 and E2 is semi-reflexive with respect to Ex. Thus <£2,£!> is reflexive if (El[Zh(E2)'])'= E2 and (E2lZb(E1)']), = E1. (1) A locally convex space E[%~\, where % is the Mackey topology, is reflexive if and only if <£',£> is a reflexive dual pair. This is a direct consequence of 5.(3) b). From 5.(3) a) we obtain (2) A dual pair <£2, £t > is reflexive if and only if Ex and E2 are weakly quasi-complete, and if and only if every bounded set in El and in E2 is weakly relatively compact. There exist semi-reflexive spaces which are not reflexive. If E[X] is reflexive, it is only necessary to replace the Mackey topology I by a strictly coarser one which still gives E' as the dual. This is of course trivial, since it results from using a topology on E which is clearly unsuitable. What is the position when £ is the Mackey topology? Put another way, are there dual pairs <E2,£i> m which only one of the two spaces is semi-reflexive with respect to the other? (3) Suppose that <£2,£i> is a dual pair. a) £i[£fc(£2)] is barrelled if and only if E2[Zk(El)'] is semi-reflexive. b) <£2,£i> is reflexive if and only if El[Xk(E2)] and E2\_<Xk(E1j] are barrelled. c) El[Zk(E2)~\ is semi-reflexive but not reflexive if and only if ^Pk(^i)] is barrelled but not semi-reflexive. Proof. To say that Ex \%k{E2)~\ is barrelled is to say that the bounded sets in E\=E2 are relatively ^(EJ-compact. Since E'2 = El, this is the same as saying that E2\_<Xk(El)~] is semi-reflexive. This proves a). b) follows directly from a). It follows from a) and b) that if Ex [<%k(E2)~] is semi-reflexive and not reflexive then £2[^Xfc(£1)] is barrelled and not semi-reflexive. Conversely if E2\_<Xk(El)~\ is barrelled and not semi-reflexive then £i[Sk(£2)] *s indeed semi-reflexive, by a), but it cannot be reflexive, for it would then be barrelled, which is not possible, by b). Thus c) is also proved. The following refinement of 1.(3) follows from (3) a) and 3.(2): (4) Suppose that £[£] is locally convex and that £ is the Mackey topology. E[X] is barrelled if and only if E' is %s(E)-quasi-complete. In another form: The locally convex space £[£] is weakly quasi-complete if and only if £'[£fc(£)] is barrelled. It follows from (3) c) that 20 Kothe, Topological Vector Spaces I
306 § 23. The bidual space. Semi-reflexivity and reflexivity (5) We obtain all the locally convex spaces with the Mackey topology which are semi-reflexive but not reflexive by forming the duals E of all spaces E\%\ which are barrelled but not semi-reflexive and giving them the Mackey topology %k(E). In particular, from § 21,5.(3) we get (6) // £[£] is a non-reflexive (F)-space, F[£fc(£)] is semi-reflexive, but not reflexive. 7. Distinguished spaces. We saw in 6. (3) a) that the semi-reflexivity of £[£] corresponds dually to E'[%k(Ej] being barrelled. On the other hand we know (3.(4)) that the strong dual E'[Zb(E)~\ of a semi-reflexive space E[X~\ is always barrelled. We shall now determine a property of E[X] which corresponds dually to the strong dual being barrelled. A locally convex space is said to be distinguished if every ^'-bounded subset Bx of the strong bidual E" lies in the %S(E\ ^-closure in E" of a bounded set B of E. This is the same as saying that for every Bx there is a B with Bx cz B°°, where B°° is the polar in E" of B° cz E'. We now have (1) A locally convex space E[X] is distinguished if and only if the strong dual F[£b(£)] is barrelled. Proof. Suppose that £[£] is distinguished. As B runs through the bounded sets of E, the sets B° form a base of £b(£)-neighbourhoods of o in E', and the sets (B00)0 = B° form a base of £b(F')-neighbourhoods of o, i.e. Zb(E) and Zb(E") are the same topology on E'. E'\%b(E)\ is therefore barrelled. Conversely if E'[%b(E)~\ is barrelled, so that Zb(E) and Zb(E") are the same, then every bounded set Bt off", being a £b(£")-equicontinuous set, lies in the polar (B°)° in E" of a £b(£)-neighbourhood B° of o, where B is bounded in E. Every semi-reflexive space is distinguished. Every (B)-space is distinguished, since the strong dual, being a (B)-space, is barrelled. A non-reflexive (B)-space therefore gives an example of a distinguished space which is not semi-reflexive. As we shall see in § 31, 7., there are (F)-spaces which are not distinguished, and whose strong duals are therefore not barrelled. (2) If E[X] is distinguished, E" is the ZS(E')-quasi-completion of E. For every bounded set Bx of E" lies in a set B°° which is £s(F)-com- pact, and so is £s(F)-complete. (3) // E\jt] is distinguished and E'[%h(Ey\ is semi-reflexive, then E' is reflexive and E" is the %k(E')-quasi-completion of E.
8. The dual of a semi-reflexive space 307 It follows directly from (1) and 5.(1) that E is reflexive. E" is the £s(F)-quasi-completion of E, by (2). Since E is the dual of E", the £s(F)-closure and the Ifc(F)-closure in E" of a bounded absolutely convex set B cz E coincide. E" is therefore the £fc(£')-quasi-closure of E (cf.§18,4.(4)). 8. The dual of a semi-reflexive space. As an example at the end of this number shows, the strong dual of a semi-reflexive space need not be quasi-complete. However, we have (1) Suppose that £[£] is semi-reflexive, and let E (respectively E) be the quasi-completion (respectively completion) of the strong dual E[Xb(EY\. Then E[%k(Ej] and E[%k(E\\ are also semi-reflexive, and their strong duals are the quasi-completion and completion, respectively, of the strong dual of £[£]. Proof. By hypothesis E\Zh{E)~\ has E as dual. By §15,9.(11), the completion E, and a fortiori the quasi-completion E, of E also have E as dual. Now the F-bounded subsets of E\%~\ are relatively £s(£')-com- pact. By § 21,4.(5), they are also relatively £s(£')-compact, and a fortiori are £s(£')-compact, and so they are also E- and E'-bounded. But by 3.(1) it follows from this that E\%k(E)\ and E\%k(E)\ are semi-reflexive. Since the subsets of E which are bounded with respect to E, E and E are the same, the strong topologies on E and on E are the restrictions of the strong topology on E. The last assertion follows from this. By making the topology of a semi-reflexive space E finer, it is therefore always possible to arrange for E to remain semi-reflexive, but for the strong dual to be quasi-complete, or indeed complete. If we make the further assumption that E is £s(£)-quasi-complete, E[Xk(Ey\ is reflexive, for both E[Xk(Ej] and E are semi-reflexive. In this case we can obtain certain information about the iterated strong duals. Thus suppose that E[X] is semi-reflexive, that £ is the Mackey topology and that F[£b(£)] is not quasi-complete, so that E[X~] is not reflexive. Further suppose that £'[£b(£)] is semi-reflexive (and so reflexive, for then £[£fc(£')] is indeed reflexive). Then E" = E, and the topology £&(£') on E" = E is strictly finer than % so that E" is strictly larger than E, although contained in E; thus £b(F) is coarser than Xb(E) = i:k(E) onE It follows by (1) from E cz E" cz E that the strong dual of E"\Zb{E)\ is again E, equipped with the topology H^E"). In this way we obtain a sequence E'czE"cz--czE and a sequence £b(F), lLh(E"),... of topologies on E which get finer and finer, and under each of which E is semi- reflexive. 20*
308 § 23. The bidual space. Semi-reflexivity and reflexivity If E is not reached after finitely many steps, we can form £^ = £'uru-. If E{0)) is also not equal to E, so that E{<0) is^lso not £b(£)-quasi-complete, then £b(£(£0)) is strictly coarser than %b(E') on E, and the method can be extended by transfinite induction, until E is reached. We give an example in which E" — E — E. It is not known if there are spaces for which a higher iterated dual is first equal to E. We consider the space cpdi the direct sum of d spaces Ea= K. Suppose that the number d is uncountable. In § 18,5., besides the topology of the locally convex direct sum we also considered the strictly coarser topology X' of the topological direct sum on cpd. By §22, 5.(5), ((pd[X'~\)' is equal to the linear subspace cod0) of cod consisting of all u = (vj e wd with only countably many non-zero va. By § 18, 5. (7), the bounded subsets of cpd [X~] and </>d[jX'] are the same, being the bounded finite- dimensional subsets. Since these are relatively 2s(cod)-compact, they are also relatively Xs (coj,0))-compact, so that </>dpX'] is semi-reflexive. The strong topology %b(cpd) on cod coincides with the weak topology Xs{cpd), and in particular the strongly bounded and weakly bounded subsets of cod coincide. The bounded subsets of cod0) are the intersections of the bounded subsets of cod with cod°\ The set B of all u = (va)ea)d with |t?J^l for all a is bounded and Xs(q>d)- compact in cod, so that B{0) = Bncod0) is bounded and weakly closed in cod0). But B{0) is 2s((/>d)-dense in B, so that B(U) is weakly precompact, but not weakly compact, in o)d°\ But this means that the 2b((/>d)-bounded subsets of cod0) are in general not relatively weakly compact; cod0) is thus not semi-reflexive. By 2.(3), (o)d0))" = a)d. If (pd is given the topology Xk((Dd0)\ which lies between X and X\ cpd is semi-reflexive; the strong dual cod0), whose topology coincides with the weak topology, is not quasi-complete, but (cpd)'" is equal to the completion (Dd of (Dd°\ The space q>d9 which is semi-reflexive, but not reflexive, under the topology Xk(o)d°\ becomes reflexive when we pass to the finer topology Xk((Dd). Komura [2] has given an example of a reflexive locally convex space which is not complete. 9. Polar reflexivity. We can raise the questions which lead to the concepts of semi-reflexivity and reflexivity for topologies other than the strong topology. By the Mackey-Arens theorem these questions are only non-trivial for topologies which are finer than the Mackey topology. We shall investigate these questions for the polar topology £°, that is, for the topology of precompact convergence. Thus we call a locally convex space E[Z~\ polar semi-reflexive if (£'p°])' = £, and polar reflexive if further Z = Z°° (cf. §21, 6. and 7.). (1) Suppose that E[X] is locally convex. The space {E[%°~\)' is equal to (J C, where (£ is the class of absolutely convex precompact subsets of E[X], and C is the completion of C in E[pL~\. £[£] is polar semi-reflexive if and only if every precompact subset of E is relatively compact.
9. Polar reflexivity 309 Proof. £[£] c= (£')*. The set C is compact, and therefore weakly compact, and so it is the weak closure of C in E'*; the first assertion therefore follows from 2.(1). The second part of the theorem follows directly from this. The class of polar semi-reflexive spaces is larger than the class of semi-reflexive spaces. (2) Every quasi-complete space E\1L\ is polar semi-reflexive. In particular, every semi-reflexive space is polar semi-reflexive. The first assertion follows from (1). If £[£] is semi-reflexive, every bounded subset of E[X] is relatively weakly compact, and a fortiori so is every precompact set, and the assertion follows from (1). Analogously, we have (3) Every reflexive space is polar reflexive. For £[£] is polar semi-reflexive and £ is the strong topology, and it was shown in § 21, 7. that this coincides with £°°. (4) // £[£] is quasi-complete, if X is the Mackey topology and if £'[£°] is also quasi-complete, then E[%~\ is polar reflexive. Proof. By (2), £[£] is polar semi-reflexive. We still have to prove that <X = <X°°, and so we must show that the absolutely convex weakly relatively compact subsets K of E' are the same as the £c-precompact subsets C of E'. Because £'pX°] is quasi-complete, every C is relatively ^-compact, and a fortiori is weakly relatively compact. Since every set C is contained in an absolutely convex set of the same kind, all the sets C are also sets K. The converse follows from § 21, 6.(2). It follows directly from (4) and § 21, 6.(4) that (5) Every (F)-space is polar reflexive. These ideas have a simple connection with the duality theory of abelian topological groups (Freundlich-Smith [1]). Clearly, with respect to addition alone, every locally convex space E[%~] is an abelian topological group. A character x on an abelian topological group G is a continuous group homomorphism from G into the multiplicative group of complex numbers of modulus 1. The character x(x) = 1 wiU be denoted by 1. The characters of G form a group G under multiplication, with 1 as unit element. We now consider the special case of a real locally convex space £[£]. We can assign a character xu t0 eacn element u of the dual space E\ by setting xu(x) = eiux. Certainly xu(xl+x2) = xu(x1). Xu{xi\ and Xu ls continuous since u and the exponential function are both continuous. We shall now show that all the characters on E[%~] are obtained in this way. (6) The correspondence u^Xu is an algebraic isomorphism of the additive group E onto the multiplicative character group Epf E[X~\. Proof. The linear functional u = 0 corresponds to the character 1. If w#0, then u does not vanish identically on any neighbourhood U ofo with \u(U)\<2n, so that eiux + 1 and the correspondence u^>xu is one-one.
310 § 24. Some results on compact and on convex sets It remains to show that every character Xo nas tne f°rm Xu- Because x0 is continuous, there is a neighbourhood U of o with \x0(x)—l\<n for all xeU. A continuous additive function is defined on U by setting u0x = — log%0(x) = arc%0(x), where the principal value is taken. It is real linear on U, for linearity with rational coefficients follows from additivity, and linearity for real coefficients follows from this by taking limits in the usual way. Now <if z is an arbitrary element of £, — zeU for suitable n. If we put /l \ . . n UqZ = yi'Uq[ — z J, u0 is extended in a continuous linear way to the whole of £, so that u0eE'. Since XoM = e^Uox holds on U9 Xo^) = eiu°z holds on the whole of E. If therefore we start from a real dual pair (E2,Exy, by setting {u,x} = eiux we obtain a dual pairing of the two abelian groups E2 and £x, under which each element of one group determines a homomorphism of the other group into the group of complex numbers of modulus 1. If we take the Mackey topology as the topology on each of the groups, then by the Mackey-Arens theorem each group is the character group of the other. The well-known Pontryagin duality theorem says that the character group G of a locally compact abelian group G is again locally compact under the topology of uniform convergence on the compact subsets of G, and that conversely G is the character group of G. Since E[%~] is only locally compact if £ is finite-dimensional, this theorem only includes the trivial case for locally convex spaces. On the other hand to every pairing {£',£} corresponding to a polar reflexive £[£], there corresponds an analogue of the Pontryagin duality theorem; in particular by (5) a real (F)-space and its X°-dual are each the character group of the other, with respect to the topology of uniform convergence on compact subsets. § 24. Some results on compact and on convex sets 1. The theorems of Smulian and Kaplansky. As in § 3, 4., a subset M of a topological space R is said to be countably compact or relatively countably compact, if every sequence in M has an adherent point in M, or in R, respectively; M is said to be sequentially compact, or relatively sequentially compact, if every sequence in M has a subsequence convergent to an element of M, or of R, respectively. Every (relatively) sequentially compact set is (relatively) countably compact. We remark that the closure of a relatively countably compact set need not be countably compact; a similar remark applies to relatively sequentially compact sets (cf. Grothendieck [6], for example). (1) If M is a weakly relatively countably compact subset of a locally convex space E[X], then M is bounded. For otherwise there would be a sequence xneM and an element ueE' with |wx|->oo, and xn could then have no weakly adherent point.
1. The theorems of Smulian and Kaplansky 311 We now give an example of a non-reflexive (B)-space in whose dual there is a weakly relatively countably compact set which is not weakly relatively sequentially compact. The space Z1 is a subspace of (/°°)'. The set of elements c1,c2,...e/1 is bounded in (/°°)', and is weakly relatively countably compact, being a subset of the unit ball of (J00)', which is weakly compact, by the Banach-Alaoglu theorem. But no subsequence e„ is weakly convergent in (I00)'; to see this, it is enough to consider the sequence ue„, where u is an element of /°° whose nrth coordinates are alternately equal to 0 and 1. We now obtain Smulian's theorem: (2) Suppose that E[X] is locally convex. If E' is weakly separable, every weakly relatively countably compact subset of E is weakly relatively sequentially compact. Proof. Let xn be a sequence in E each of whose subsequences has at least one weakly adherent point in E. We must show that xn has at least one subsequence which is weakly convergent in E. By hypothesis there is a weakly dense sequence um in E'. Since the sequence xn is bounded, using a diagonal procedure we can find a subsequence xnu=yk, for which \im umyk exists, for each m. The sequence yk has a weakly k->oo adherent point y in E. If z is any weakly adherent point of the sequence yk, then it follows from umy = lim umyk = umz that um(y — z) = 0 for each um9 k-* oo and the um are dense in E'. Thus y = z, and the sequence yk has only one weakly adherent point. But it follows from this that y is the weak limit of the yk: for if an infinite subsequence ykj were not to lie in a weak neighbourhood U(y), ykj would have a weakly adherent point which would have to be different from y. It follows directly from §21, 3.(5) that Smulian's theorem holds for every separable metrizable locally convex space. The following generalization of (2), due to Dieudonne and Schwartz [1], is substantially stronger: (3) Suppose that E[Z~\ is locally convex. If there is a metrizable locally convex topology %' on E which is coarser than X, then every weakly relatively countably compact subset of E is weakly relatively sequentially compact. In particular this holds for every metrizable locally convex space. Proof. Since X is metrizable, there is a sequence Ux => U2 => * * * of 00 absolutely convex ^'-neighbourhoods of o in E, with f] Un = o. By n=i hypothesis the Un are also ^-neighbourhoods of o. Suppose that xn is a sequence in E with the property that every subsequence has a weakly adherent point, and let H be the closed linear span of the xn in E. Then the set of xn is total in H, and the sets Vn=UnnH form a sequence of
312 § 24. Some results on compact and on convex sets ^-neighbourhoods of o in H, with f] Vn = o. Thus [j V°n is weakly m= 1 n— 1 dense in H'. Using exactly the same argument as in the proof of § 21, 3.(5), it follows that H' is weakly separable. Since all the weakly adherent points of the subsequence of xn lie in H, xn has a £s(//')-convergent subsequence x by Smulian's theorem. But by § 22,2.(1), the topologies £s(/f') and £S(F) coincide on H, so that xn is weakly convergent in E, and consequently the set {xn} is weakly relatively sequentially compact. We observe that the hypothesis of (3) is more general than the hypothesis of (2). For if AT is a countable weakly dense subset of E, then by §21,1.(2) £S(N) is a metrizable locally convex topology on E, and it is coarser than % by § 20, 2. (4). (4) Every weakly relatively countably compact subset of a strict (LF)-space is also weakly relatively sequentially compact. For the weak closure of a weakly relatively countably compact set M 00 is bounded, every bounded subset of a strict (LF)-space £ [£] = £ £„[£„] M=l always lies in an (F)-space En, by §19,4.(4), and the weak topology on En coincides with the weak topology of E. The assertion therefore follows from (3). (5) Suppose that E[%] is locally convex and that it satisfies the hypotheses of (2) or (3). // xn is a sequence in E each of whose subsequences has at least one weakly adherent point in E, then each of these adherent points is also the weak limit of a suitable subsequence. Thus the weakly sequentially compact and the weakly countably compact subsets of E are the same. Proof. If y is a weakly adherent point of xn then, in the case where the hypotheses of (2) hold, a diagonal procedure produces a subsequence xnk with umxnk-+umy, and as before it can be shown that y is the weak limit of the sequence xnr Case (3) can again be reduced to case (2). We now obtain the following theorem of Kaplansky's (cf. Bourbaki [6], Vol. 2, p. 82) (6) Suppose that E\%~\ is locally convex and that E is the union of countably many weakly compact subsets. If M is an arbitrary subset of E, every weak closure point x0 of M is always a weak closure point of a countable subset of M. Proof. Suppose that E is the union of the weakly compact sets Cj a C2 cz •••. We consider a weak neighbourhood of x0 of the form
2. Eberlein's theorem 313 |wt-(x0 — x)\ < —, i=l,..., k, with/c elements u{eCn. By hypothesis there m is a yeM in this neighbourhood. Since x0 and 3; are weakly continuous on E\ and so also on C„, there are weak neighbourhoods V{ of the uh 1 for which |^-(x0—y)\ < — holds for all i^eVJ. We consider such a m system of neighbourhoods for each/c-tuple ul,...,uk and a corresponding yeM. By Tychonoff's theorem the /c-times topological product C* is compact, so that C\ is covered by finitely many Vx x ••• x Vk; there is therefore a finite subset Mnkm of M with the property that at least one element of Mn k m lies in each neighbourhood of the form |wf(x0 — x)\ < —, ' ' 00 m i=l,...,k,uieCn. The set (J Mnkm has the required properties. n,k,m= 1 By §20,7.(6), every convex weakly sequentially closed subset of a metrizable locally convex space is always weakly closed. The theorems of Smulian and Kaplansky enable us to make another assertion in this direction: (7) Suppose that E[Z~\ is a metrizable locally convex space or a strict (LF)-space. If x0 is a weak closure point of a weakly relatively compact subset M of E, x0 is the weak limit of a sequence in M. A subset of E is therefore weakly compact if and only if it is weakly relatively compact and weakly sequentially closed. Proof. If E[X] is metrizable, and if Ul => U2 => • * * is a base of neighbourhoods of o for E[X~], then E' is the union of the weakly compact sets U°n. By (6), x0 is therefore an adherent point of a sequence in M, and the assertion follows from (5). The case of an (LF)-space can be reduced to this one as in (4). 2. Eberlein's theorem. We have already given an example in § 3,4. of a sequentially compact (and therefore countably compact) set which is not compact. We can also describe this example in the following way: in If, with d>N0, let H be the subspace consisting of all vectors x = (<^a) with at most countably many non-zero coordinates £a; then the unit ball of H is £s(/j)-sequentially compact, but is not £s(/j)-compact. The following theorem of Eberlein's, proved first by Eberlein [2] for (B)-spaces, and later generalized by Grothendieck [6], gives a far-reaching condition for the identity of relatively countably compact and relatively compact sets: (1) // the locally convex space £[£] is quasi-complete under the Mackey topology ^k(Ef), every relatively countably %-compact subset M of E is relatively compact.
314 § 24. Some results on compact and on convex sets Indeed, we have (!') If £pX] is locally convex, if M is relatively countably compact and if C (M) is ^-complete, then M is relatively compact. Proof, a) First we show that it is sufficient to prove (1') under the hypothesis that £[£] is ^-complete. Suppose therefore that M is a relatively countably compact subset of £[£]. We form the ^-completion E{Xk~\ of^£. By §21,4.(5), %k is the Mackey topology on E. The extension of £ to E will again be denoted by %. M is then also relatively countably compact in E[%~]. We assume that it has been proved that M is relatively compact in E; we must then show that M is already relatively compact in E. For this it is enough to show that the closure M of M in E is a subset of E. But the closed convex cover of M in E is ^-complete, by hypothesis, and so it is also ^-closed in E. As it is a convex set, it is also ^-closed in E, and consequently it contains M. b) By § 5, 6.(3) every relatively countably compact set is precompact. It is therefore sufficient to show that the closure M of M in E[X~\ is complete, since a complete precompact set is compact. We may also assume that £ is the weak topology. For M is also relatively countably compact with respect to the weak topology, and if the weak closure of M is weakly complete, then by § 18,4.(4) the ^-closure of M is also ^-complete. Consequently (V) is reduced to the following assertion: (2) Suppose that E\%\ is Uncomplete, and that M is a weakly relatively countably compact subset of E. Then every weak closure point z of M in (E')* always belongs to E. Proof. By Grothendieck's theorem (§21,9.(4)) it is sufficient to show that z is weakly continuous on every absolutely convex weakly compact subset K of E'. We suppose that this is not the case, so that there exists a K on which z is not weakly continuous. By §21, 6.(5), z is not continuous at o on K, so that there is an e > 0 with the property that in each weak neighbourhood of o there is a ueK with |wz|^e. We now inductively construct two sequences uneK, xkeM for which the following inequalities hold: (3) \UiXn-UiZ\^-, l^i^n-l, n (4) k*<l^-, l^i^n, n (5) |«„Z| ^ 8.
3. Further criteria for weak compactness 315 For if xl9...9xn_l9 ul9...9un_l have already been constructed, we determine xneM in such a way that (3) is satisfied. This is possible, since z is a weak closure point of M. We then determine un in such a way that (4) and (5) are satisfied. This is possible because z is not continuous at o. The sequence xneM now has a weakly adherent point x0 in £, and likewise the sequence uneK has a weakly adherent point u0eK. As far as these adherent points are concerned, it follows from (3), by letting n tend to infinity, that uix0 = uiz9 and likewise it follows from (4) that 0 i = 0. As x0 is an adherent point of the sequence xn9 it follows that u0x0 = Q. But x0eE is weakly continuous on K9 so that uoxo = 0 is adherent to the sequence of values i 0 — i z; this contradicts (5). Thus (2) and (1') are established. New criteria for reflexivity follow directly from Eberlein's theorem and theorems § 23, 3.(1), (3) and § 23, 5.(4): (6) A Xk-quasi-complete locally convex space is semi-reflexive if and only if every bounded subset is weakly relatively countably compact. (7) An (F)-space is reflexive if and only if every bounded subset is weakly relatively countably compact. (8) A (B)-space is reflexive if and only if its closed unit ball is weakly relatively countably compact. Remark. BySMULiAN's theorem, 1.(3), "weakly countably compact" can be replaced in (7) and (8) by "weakly sequentially compact". It follows directly from 2.(1) and 1.(7) that (9) // M is a weakly relatively countably compact subset of an (¥)- space, the set of all the limits of weakly convergent sequences in M is weakly compact. By Eberlein's theorem, the weakly closed weakly countably compact subsets of a ifc-quasi-complete locally convex space are the same as the weakly compact sets. A fortiori this also holds for the weakly closed Xa-compact sets (cf. § 3, 4.). 3. Further criteria for weak compactness. A subset M of a Hausdorff topological space R is said to be pseudo-compact if every continuous real-valued function f(x) on M is bounded. M is said to be relatively pseudo-compact if for each unbounded continuous function/ on M there exists a point x0 in the closure M, in all of whose neighbourhoods f(x) is unbounded. The closure of a relatively pseudo-compact set is pseudo-compact. (1) Every weakly relatively pseudo-compact subset M of a locally convex space E[%] is bounded. Every weakly (relatively) countably compact set N is weakly (relatively) pseudo-compact.
316 § 24. Some results on compact and on convex sets Proof. If M is unbounded, there is an element ueE with the property that \ux\ is unbounded on M. However \ux\ is bounded on the weak neighbourhood x+Un;E of any xeE. If f(x) is an unbounded function on N, and if \f(xn)\^n on the sequence xneN, then / is unbounded on every weak neighbourhood of a weak accumulation point x0 of the sequence xn. We now have the following lemma of Ptak [3] : (2) // M is a weakly (relatively) pseudo-compact subset of a locally convex space E[X~], if z is a weak closure point of M in Er* and if ut is a sequence of elements of E\ then there exists a point x0 in M (in M) with uiz = uix0 for all i=l,2,.... Proof. The functions f(x)=\ui(x — z)\ are weakly continuous on the weak closure M of M, and they are each bounded, by (1). Suppose 1 that \f(x)\^ki on M. The function f(x) = £ -•—fi(x) is the limit of a i=\ 2lkt uniformly convergent series of continuous functions, and it is therefore continuous on M. Since we can make l/^xJI^e,..., |/w(x)|^e by choosing x in a suitable way, inf/(x) = 0 on M. If f(x) were strictly positive on the whole of M, l/f(x) = g(x) would be continuous_and unbounded on M. But then by hypothesis there would be an x0eM in all of whose weak neighbourhoods g(x) would be unbounded. Then f(x) would have infimum 0 in each of these neighbourhoods, so that, since/is continuous at x0, f(x0) would have to be zero. But this means that uix0 = uiz for all i. A subset M of a locally convex space E[%~] is said to be weakly (relatively) convex-compact if the following holds: suppose that Kx 3 K2 =3 • • • is a sequence of closed convex subsets of E for which all the intersections KnnM are non-empty; then the sequence KnnM has a weakly adherent point in M (in E). This concept is due to Smulian [1], [3]. (3) Every weakly relatively convex-compact subset M of a locally convex space E \%~\ is bounded. Every weakly (relatively) countably compact set N is weakly (relatively) convex-compact. If M were unbounded, there would be a point u0eE' and a sequence xteM with |w0xf|^i. The sequence K1^K2^'" of closed convex covers Kn of the sets {xn9 xn+1,...} could then have no weakly adherent point, since \u0y\^n for all yeKn. If KjD^d- and if KnnN is non-empty for all n, then a sequence xneKnnN has a weakly adherent point in N (respectively E).
3. Further criteria for weak compactness 317 (4) // M is a weakly (relatively) convex-compact subset of a locally convex space £[£], if z is a weak closure point of M in £'*, and if uv is a sequence of elements of £, then there is an x0 in M fin M) with \imui(z — xo) = 0. i-*oo _ Here M means the weak closure of M. To prove this, consider the sequence of closed convex weak neighbourhoods Un(z) defined by the 1 inequalities |Mf(z —x)| ^ —, for i= 1,..., n. A weakly adherent point x0 n of the sequence Un(z)nM has the required properties. A bounded subset M of a locally convex space E[%~] is said to be weakly (relatively) partially compact if the following holds: if a sequence of elements of M has a weakly adherent point z in £'*, then for each sequence ut contained in an absolutely convex weakly compact subset of E there exists an x0 in M (in M) with \imui(z — xo) = 0. This concept was introduced by Day [8]. I_+0° (5) Suppose that M is a subset of the locally convex space £[£]• // M is weakly (relatively) countably compact or weakly (relatively) pseudo-compact or weakly (relatively) convex-compact, then M is also weakly (relatively) partially compact. This follows from (1), (2) and (4). Eberlein's theorem can be strengthened in the following way: (6) Suppose that £[£] is a locally convex space and that M is a subset whose closed convex cover C (M) is Uncomplete. M is weakly relatively compact if and only if it is weakly relatively partially compact. The proof proceeds analogously to that of 2.(1): If M is weakly relatively compact then it follows from (5) that M is weakly relatively partially compact. Suppose conversely that M is weakly relatively partially compact, and that z is a weak closure point of M in (£')*. It is sufficient to show that z belongs to the ^-completion E of £, for then z, being a weak closure point of M, lies in C (M), which is weakly closed in £, since it is ^-complete. We proceed as in 2.(2) and suppose that z is not weakly continuous on some absolutely convex weakly compact subset K of E. Then again there are sequences ukeK and xkeM for which the inequalities (3), (4) and (5) of No. 2 hold. Since M is bounded, it is weakly precompact, so that the sequence xneM has a weakly adherent point 3c0 in (F)*, and likewise uneK has a weakly adherent point u0eK. As in the proof of 2.(2) we obtain the relations uix0 = uiz, uoXi = 0 and uoxo = 0. Applying the hypothesis that M is weakly relatively partially compact to the sequence wl9 w0, w2,
318 § 24. Some results on compact and on convex sets u0,... in K and to x0, it follows that there is an x0 e M with lim ut(x0 — x0) = 0 and with uoxo = uoxo = 0. But uoxo = 0 is adherent to the sequence of values utx0, so that 0 is a closure point of the values uix0 = uiz; this contradicts 2.(5). Consequently z is weakly continuous on each K, so that it belongs to E, by Grothendieck's theorem. (7) Suppose that M is a subset of a locally convex space E[X], and that the closed convex cover C (M) is Uncomplete. In particular this hypothesis is satisfied by any bounded subset of a Xk-quasi-complete space. Then the following properties of M are equivalent, a) weak relatively countable compactness, b) weak relative pseudo-compactness, c) weak relative convex-compactness, d) weak relative partial compactness and e) weak relative compactness. By (5) and (6), e) follows from each of the properties a) to d). Conversely a) and d) follows from e); b) then follows from a), by (1), and c) follows from a), by (3). The equivalence of weak relative compactness of M with b) was proved by Ptak [2], [3], the equivalence with c) byDiEUDONNE[ll] and the equivalence with d) by Day [8]. Further criteria for weak compactness are given in Day's book [8] and the work of Grothendieck [6]; we shall consider one of these in No. 6. Applying 1.(3), we obtain the following special case of (7): (8) The following properties of a subset M of an (F)-space are equivalent: a) weak relative countable compactness, b) weak relative pseudo- compactness, c) weak relative convex-compactness, d) weak relative partial compactness, e) weak relative compactness and f) weak relative sequential compactness. As we saw at the beginning of No. 2, a weakly countably compact set need not be weakly compact; the assertion of (7) is therefore not true in general if the word 'relative' is omitted throughout. For (F)- spaces, however, we have (9) In an (F)-space E the weakly compact subsets are the same as the weakly countably compact sets, the weakly sequentially compact sets and the weakly convex-compact sets. First we show that a weakly convex-compact subset M of E is weakly sequentially closed. Suppose that xneM is weakly convergent to x0, and let H be the closed linear subspace of E defined by the xn. Since H is separable, there is a weakly dense countable set N in H', by § 21, 3.(5). The topology %S{N) on H is therefore Hausdorff and metrizable. There is therefore a sequence Kx^> K2^> •- of closed absolutely convex
4. Convex sets in spaces which are not semi-reflexive. The theorems of Klee 319 00 £s(N)-neighbourhoods of x0 in H with f] Kn={x0}. Since KnnM is n = 1 non-empty for each n (indeed it contains infinitely many xk) x0 belongs to M, which is assumed to be weakly convex-compact. By (8) a set M which satisfies one of the conditions of (9) is weakly relatively compact. Since it is also weakly sequentially closed, it follows from 1.(7) that it is weakly compact. (8) and (9) also hold for strict (LF)-spaces. The assertion corresponding to (8) follows from (7) and 1.(4), and the one corresponding to (9) from (9) and § 19, 5.(4). By § 23, 3.(1), new criteria for reflexivity are given by (7) and (8); let us state the one produced by the equivalence of weak relative compactness and property c): (10) Suppose that £[£] is locally convex and Xk-quasi-complete (respectively an (F)-space). E is semi-reflexive (respectively reflexive) if and only if every decreasing sequence of closed bounded non-empty convex subsets of E has a non-empty intersection. 4. Convex sets in spaces which are not semi-reflexive. The theorems of Klef:. If a £fc-quasi-complete space E[X] is not semi-reflexive, then by 2.(6) there is a bounded sequence in E with no weakly adherent point. This sequence defines a closed separable linear subspace of £[£] which cannot be semi-reflexive. Thus we obtain the following criterion for reflexivity from 2.(6): (1) A ^-quasi-complete locally convex space is semi-reflexive if and only if every closed separable linear subspace is semi-reflexive. If E is not semi-reflexive and not separable, E therefore contains a closed linear subspace of infinite codimension in E which is also not semi-reflexive. We now show that the same also holds in the separable case: (2) // E[%~\ is a ^-quasi-complete locally convex space which is not semi-reflexive, then there exists a closed linear subspace of infinite co- dimension which is also not semi-reflexive. Proof. By 2.(6), E contains a bounded sequence x{n) with no weakly adherent point. We form the closed linear span H of the x(n) in E. If B is the closed absolutely convex cover of the x(n) in £, H also contains oo the normed space EB= \J nB; the norm topology on EB is finer than n= 1 the induced topology, by §20,11. Let z1,z2,... be a sequence of linearly independent elements of EB c H. We set Fk = [zj,..., zfe]. The sequence
320 § 24. Some results on compact and on convex sets Fk in H' is decreasing, and each Fk has codimension 1 in Fk_ l. Corresponding to the spaces Fk we can find an increasing sequence of algebraic complements Gk in H\ and each Gk has dimension k. By § 15, 8.(2), the decomposition H' = Fk® Gk is ^-complementary. Then by § 20, 5. (1) and §15,8. (2) the spaces Hk = Gk form a sequence of topological complements to the spaces Fx c= F2 <= •••, with Hx => H2 => ••*, #fc of codimension 1 in Hk_u and Hx of codimension 1 in H. Each x(n) then has the unique representation x<»> = <*»>Zl + ••• + ^^x^, Jc^eHfcn^. Using a diagonal procedure, we now choose a subsequence of the x(n\ again denoted by x(r°, with the property that lim £,kn) = yk exists for all k. n-* oc Then for each k we can determine an nk and a .y(k) = yiZi + ••• + ykzk + 7/c + i zk + i + £fe+k)i ^ £B with yic + i + 7/c + i in sucn a waY tnat tne ^B-norm distance between y{k) and x("k) satisfies ||yk) —x(Wk)||^l/fc. It follows from 7fe+i+7fe+i tnat y{l) — y{k\ ..., y(k~l) — y{k) are linearly independent, and so therefore are y{1\...,yik~1}, modulo Hk. It follows from this that the closed linear subspace L spanned by yi2\ y{4\... has infinite codimension in H, and thus in E as well. Since ||y2fc) —x("2k)||^ 1/2/c, the yi2k) form a bounded sequence in EB, and a fortiori in L. A point weakly adherent to the sequence yi2k) must also be weakly adherent to x("2k). From this it follows that L is not semi-reflexive. We now give a further criterion for reflexivity: (3) A %k-quasi-complete locally convex space E[X] is semi-reflexive if and only if every bounded closed absolutely convex subset K has a supporting hyperplane parallel to each real closed hyperplane H. Proof. It is sufficient to prove the theorem for real E (cf. § 21,11.(2)). a) Suppose that E is semi-reflexive. Then K is weakly compact, and the existence of the supporting hyperplane follows from § 20, 7. (8). b) Suppose that E is not semi-reflexive. Given any hyperplane Hbo, we shall produce a bounded closed absolutely convex set which has no supporting hyperplane parallel to H. By §23,3.(6), H is not semi-reflexive. By 3.(10) there is a decreasing sequence of bounded non-empty convex ^-closed subsets C„ of H with empty intersection. Suppose that x0 is an element of E not lying in H. We form the closed absolutely convex cover K of the sets I 1 )x0 + C„. Then the only possible support hyperplanes parallel to H are x0 + H and — x0 + H. Because K is symmetric about o it is sufficient to show that Kn(x0 + #) is empty.
4. Convex sets in spaces which are not semi-reflexive. Th theorems of Klee 321 Let x0 + y be an element of x0 + H. There exists a Cko with y$Cko and /c0^2. Since Cko is closed, there is an absolutely convex ^-neighbourhood U ofo in H with y$Cko+U. In order to show that x0 + y$K9 it is sufficient to show that there is a ^-neighbourhood Kofo in E for which x0 + y+V contains no element of the form K0= P 111 I x0 + C„ The elements of K0 have the form z = (l— p)x0 + z' with 1— p m / i \ m m = Ea«(1 and z'= Ea«:Vn' y«eC»i> EW^1- We determine a (5>0 for which 5Cl a — holds. We can then determine a p(<5)>0 in such a way that z = (l—p)x0 + z\ with p^p(3), only lies in K0 if in m the expression z' = £a„.yn, the sum £ |aw| over ocl,..^<xko-1 and any i subsequent negative ocn has magnitude less than (5, and the remaining sum Yj^n lies between 1 —5 and 1. It is easy to confirm that we can 5 choose a number < for p(d). K-\ Now let V be the neighbourhood of o in E consisting of all ax0 + x, U |a|^p(<5), xe—cz H. Elements (1 — p)x0 + z' of K0 only lie in x0 + y+V if pf^p(S). But then z has the form z = £ ain> yn> + £ a„» y„» with I«„'^^Qcy and E <V )V G c*o + y> so that ^eQ0 + fl/. However all the elements of x0 + y + K have components in H which 1/ U lie in yH . Therefore z'ey-\ must hold, which contradicts 3 3 y$Cfco + U. Thus K0n(x0 + },+ K) is empty, and so (3) is established. We now use this example to give some further constructions. Let k be chosen so that o$Q. If K is our set, constructed as above, and M is the closed convex set x0 —Cfc, then we assert that C(Ku M) is not closed. To show this, we prove that x0$ C (K u M), but that x0 is a closure point of C (K u M). We have already shown that every element of K has the form (1— p)x0 + z\ with p>0 and z in H. An element of C(XuM) therefore has the form x=[a1(l-p) + (l-a1)]x0 + a1z/-(l-a1)z//, z'etf, z"eCk, O^a^l. Because p>0 and x0$H, x could be equal to x0 only if <x1=0; in this case, though, xeM, and so it is different from x0. On the other 21 Kothe, Topological Vector Spaces I
322 § 24. Some results on compact and on convex sets hand, however, if n^k and )/„eC„cCk, (1 )x0 + yneK and x0 — yneM, so that i —)x0+yn n + \(*o-y») = (l -^)x0eC(KuM), hence x0 is in the closure of C {K u M). In exactly the same way it can be proved that K + M is not closed, since 2x0 is not in K + M, but belongs to the closure of K + M. Applying § 20, 6.(5) and § 15, 6.(10), we obtain (4) The following two conditions are each necessary and sufficient for a ^-quasi-complete locally convex space to be semi-reflexive: a) the convex cover of two bounded closed convex sets is always closed. b) the sum of two bounded closed convex sets is always closed. (3) and (4) were proved for (B)-spaces by Klee [1]; the following theorem can also be found, for (B)-spaces, in Klee [2] II. (5) A Xk-quasi-complete locally convex space E[%~\ is semi-reflexive if and only if every two disjoint bounded closed convex sets can be separated by a closed real hyperplane. Proof, a) It follows from the separation theorem for compact sets (§20,7.(1)) that in semi-reflexive spaces it is even possible to separate the sets strictly. We must therefore construct a counterexample in any non-semi-reflexive Xfc-quasi-complete space E[%~\. b) Once more we can take £[£] to be real and not semi-reflexive, and can take X = Xk. Using the ideas of the proof of (2), we can find a bounded sequence y(k) in E with the following properties: the y{2k) define a closed linear subspace L of infinite codimension in £, and have no weakly adherent point in L, while the yi2k+1) are linearly independent modulo L. Let H be the separable closed linear subspace of E defined by all the y{k). Let B be the closed absolutely convex cover of the y{k\ 00 and let xn be a X-dense sequence in B. The space HB= [j nB<=H n= 1 with B as unit ball is a (B)-space, since E[X] is quasi-complete, and its norm topology is finer than X L has infinite codimension in H, and is not semi-reflexive. We shall now construct the required counterexample in H. c) The set formed by o and the sequence xjn is compact in HB and, by § 20, 6.(3), so also is its closed absolutely convex cover A in HB. Since A is also ^-compact, the set K0 = A + L is absolutely convex and ^-closed, by § 15, 6.(10), and so also is Kx=K0nB cz HB. d) We show that o is a boundary point of K'0 = K0nHB in HB. To do this, we form the closed linear subspace L(n) of H determined by
5. Krein's theorem 323 L and xl,...,xn; L has finite co-dimension in this space, by §15,5.(3). Thus HB/L(n)nHB is infinite-dimensional. Since the image K'0 of K'0 under the canonical mapping of HB onto HJL(n) n HB lies in B, n+\ there is an element yneHB~K'0 whose HB-norm satisfies \\yn\\ <—. n Thus we have established the existence of a sequence in HB~K'0 which converges to o. e) There is no X-closed supporting hyperplane of Kx through o. Such a hyperplane must take the form ux = 0, with ueH'. But since Kt contains both xjn and — xjn, we must have uxn = 0 for all n. But the set of all xn is total in //[I], and so it follows that u = 0. 0 There is an x0eHB for which the ray px0, p>0, is disjoint from K0. If there were no such ray, K'0 would be absorbent about o; it would therefore be a barrel in HB, and so by §21,2.(1) it would be a neighbourhood of o in HB, contradicting d). Since K0 contains L, x0 does not lie in L. g) We now construct the second closed convex set K2. Since Bx=LnB is not weakly compact, by 3. (7) there is a decreasing sequence Cn of bounded convex ^-closed subsets of ^B{ which have an empty intersection. We set K2 equal to the closed convex cover of the sets — + Cn. As in the proof of (3) we see that K2 only contains elements n of the form px0 + z, with p>0, zeL. By 0, Kx and K2 are disjoint. h) Suppose that ux = y is a closed hyperplane in H which separates Kx and K2; suppose that Kx lies in ux^y. Since ogK15 y^O. Let y ux0 = <x. It follows from Cn c= \BX and Bx cz Kx that u(Cn) ^ —. Then letting rc-» oo in u\ — x0 + Cn) = h u(Cn) ^ y, it follows that y = 0. \n J n The hyperplane separating Kx and K2 must therefore pass through o, contradicting e). 5. Krein's theorem. In what follows we shall need certain results about the space C(K) of continuous functions on a compact topological space K. By § 14, 9., C(K) is a (B)-space under the norm ||/|| = sup|/(x)|. xeK By the topology £p of pointwise convergence on C(K) we mean the locally convex topology defined by the neighbourhoods sup |/(xf) —/"oW^e, x(eK. We can relate this topology in a simple i=l n way to the weak topology on C(K). Each x0eK determines a continuous linear functional dxo{f) = f{x0) on C(K). To each x0eK there 21*
324 § 24. Some results on compact and on convex sets therefore corresponds an element of C(K)' = Wl(K), which we call the point measure 5X0 corresponding to x0. This embedding K of K in 9Jl(X) is a homeomorphic mapping, when Wi(K) is given the weak topology: the weak topology on K corresponds to the coarsest Hausdorff topology on K for which all the feC(K) are continuous; the weak topology on K is coarser than the one corresponding to the topology of K\ since K is compact, it must therefore coincide with it. If we denote the linear span of K in Wl(K) by H, then Xp is equal to 2S(H). Since H c 9Jl(X), it is coarser than the weak topology %s(3Jl{K)) on C(K). We observe further that K is weakly total in yJi(K). (1) Every relatively countably Xp-compact subset of C(K) is relatively sequentially %p-compact. Proof. To prove this theorem, we reduce it to an application of Smulian's theorem. It is sufficient to show that a relatively countably ^-compact sequence fneC(K) has a ^-convergent subsequence. Let L be the ^-closed linear span of the /„ in C(K). Then {/„} is also relatively countably compact in L[2p], and L[2p] is separable. By §22,1.(1), the dual of L[Xp~] is the quotient space H/L1, where H is the linear span of K. The image K of K under the canonical mapping from H onto H/L1 is weakly compact, by §22,2.(4). If JV = {/„}, N is total in L[Xp], so that ZS(N) is a Hausdorff topology on K. K is therefore also £s(N)-compact. Since N is countable, K is therefore metrizable under the topology XS(N) = XS(L\ and so it is separable (§4,5.(2)); consequently H/LL = L! is weakly separable. Since Xp = 'Zs(H)i the assertion follows by applying Smulian's theorem (1.(2)) to fneL[Xp~]. (2) A sequence fneC(K) converges weakly to f0eC(K) if and only if the fn are uniformly bounded and converge pointwise to f0. a) Necessity. It follows directly from dX0(f„)-+5X0{f0) that /„ converges pointwise to f0. Since the sequence /„ is a strongly bounded subset of C(K\ it must also be uniformly bounded on X, so that the /„ are uniformly bounded. b) Sufficiency. We use the following theorem of Lebesgue: let \i be a positive measure on K and let L\ be the space of absolutely integrable functions with norm ||/|| = \\j \dji. If the sequence hneL\ converges /i-almost everywhere to h0 and if \hn\^g /^-almost everywhere, for some g in L*, then h0eL\ and \\hn — ho\dfi-^0. Since C(K) is a subspace of every L\, the weak convergence of /„ follows for every positive \i, and so for every measure on K. For Lebesgue's theorem cf. Bourbaki [7] Vol. 1, p. 140, for example. The following theorem, proved by Grothendieck [6], follows from (1) and (2):
5. Krein's theorem 325 (3) A subset M of C(K) is weakly relatively compact if and only if it is bounded and relatively Up-compact. Proof. The condition is necessary, for a relatively weakly compact set is also relatively compact under the coarser topology %p. On the other hand suppose that M is bounded and relatively 3p-compact. In order to show that M is relatively weakly compact it is sufficient, by Eberlein's theorem, to show that every sequence /„ in M contains a weakly convergent subsequence. By (1) /„ has a 3p-convergent subsequence, and this converges weakly, by (2). In §20,6.(3) we showed that the closed convex (respectively ab- solutely convex) cover C (M) (|~~(M)) of a compact set M is compact if \~{M) is complete. We now obtain the following sharper result, Krein's theorem: (4) The closed convex (respectively absolutely convex) cover C (K) (\~(Kj) of a compact subset K of the locally convex space £[3f] is compact if and only if C (K) (rffl) is Uncomplete. Thus if E\X\ is 3fc-quasi-complete, the closed absolutely convex cover of a compact set is always compact. (4) is clearly equivalent to (4') // K is a weakly compact subset of a locally convex space, C (K) (respectively \~(Kj) is weakly compact if and only if C(K) (P(K)) is Uncomplete. (4') was proved by M. Krein [1] for weakly sequentially compact subsets of a separable (B)-space; Phillips [2] proved (4') for (B)-spaces, and the general case, and the proof given here, are due to Grothendieck. Proof, a) First we show that it is sufficient to prove the theorem for 3^-complete spaces. Suppose that K is weakly compact in E[Z], and that C (K) (respectively r(K)) is ^-complete. We can suppose that Z = Zk. By §21,4.(5), the topology 3fc of the completion £[£&] is the Mackey topology on E. Since E and E have the same dual, K is also weakly compact in E. Because of the assumption that C (K) (respectively \~(K)) is 3^-complete, C (K) (respectively \~(K)) is also equal to the closed convex (respectively absolutely convex) cover of K in E. If we assume that (4') has been proved for £[3^], then it follows that C (K) (respectively \~{K)) is weakly compact in £, and consequently in E. b) Now suppose that K is a weakly compact subset of the complete space £[2fc]. It is sufficient to show that \~(K) is weakly compact. If Au = u is the restriction of an element ueE' to K, then A is a linear mapping from E' into C(K). Since all ue(\~(K))° have a restriction u
326 § 24. Some results on compact and on convex sets with ||m||^1 in C(K\ A is continuous for the strong topology on E and the norm topology on C{K). Now %S(E) is finer than %S(K) on E, so that A is also a continuous mapping from E[XS(E)] into C(l<£) |jlp]. If B is a closed 2fc-equicontinuous subset of £', and thus a strongly bounded 2s(£)-compact set, it therefore follows that A(B) is a bounded 2p-compact subset of C(K). It follows from (3) that A(B) is weakly compact in C(K). The adjoint mapping A' maps C(J£)' = 9W(1C) into E". A neighbourhood of o for the natural topology %n{E) on E" has the form B°, where £ is closed and 2fc-equicontinuous in E (cf. §23,4.). By the preceding argument, (A(B))° is a ^-neighbourhood of o in 9W(K). Since it follows from ue(A(B))° that A'ueB°, A' is continuous for the ^-topology on yJl(K) and the natural topology on E", and this coincides on £ with 2 = 2fc, by §23,4.(2). It follows from (A'Sxo)u = 5xo(Au) = ux0 that A' maps the set K of point measures onto K c £". K is weakly total in 9W(X), and is therefore Jfc-total as well. The image A'(yJi(K)) therefore lies in the Jn-complete linear span of K in E\ and so lies in £, since E is ^-complete. Now K lies in the weakly compact unit ball C of yJi(K); the mapping A', which is weakly continuous, by §20,4.(6), maps C into an absolutely convex weakly compact subset of E, which contains \~(K). \~(K) is therefore weakly compact. 6. Ptak's theorem. We now give a second proof of Krein's theorem. This does not use the theory of integration; instead it uses Eberlein's theorem and a combinatorial theorem of Ptak. We say that a subset M of a locally convex space £[3f] has interchangeable double limits if for every sequence xteM and every sequence UjSN, with N absolutely convex and weakly compact in E\ the existence of both the double limits \imlimUjXt and limlimlicensures that they are equal. J J We now have the following variant of Eberlein's theorem, due to Grothendieck [6] : (1) Suppose that £[2] is Uncomplete. A bounded subset M of E is weakly relatively compact if and only if it has interchangeable double limits. Proof, a) Suppose that M is weakly relatively countably compact, that xteM and UjeN, with N absolutely convex and weakly compact, and that a = \im\imujxi and j8 = limlimM-xf both exist. Let x0 be i j J i weakly adherent to the sequence xt and let u0 be weakly adherent to the sequence Uy Then MmUjX^UQXi and UmUjXi = UjX0. It then follows j i from lim w0 Xf = w0 x0 and limw-x0 = w0x0 that a = /?.
6. Ptak's theorem 327 b) Suppose that M is bounded and has interchangeable double limits. If M were not weakly relatively compact there would be a weak closure point z of M in £'* which does not lie in E. As in the proof of 2.(2), this implies the existence of two sequences xteM and ukeN, where N is absolutely convex and weakly compact in £', with properties 2.(3), 2.(4) and 2.(5). It follows from 2.(3) that \imujX~UjZ for all;, i and it follows from 2.(4) that \imu,oCj = 0, so that limlimi^oc^O. J i J The sequence UjZ need have no limit, but since MapN°, p>0, the moduli remain bounded. We can therefore choose a subsequence, denoted by u-. again, with the property that \imu:Z = P exists. It then j follows from 2.(5) that |j8|^£, so that limlimM:X~j8=t=0 = limlimMjxf. j i i J This contradicts the fact that M has interchangeable double limits. Suppose now that A = {a} is an infinite set, and that cp(A) is the space of coordinate vectors £ = (£J with only finitely many non-zero £a. We denote by C(A) the set of all \ = (Aa)e(p(A) with Aa^0, £ Aa=l. ote A If B c A, C(B) denotes the set of all I in C(A) with /a = 0 for a$B. Let © be a collection of finite subsets T of A. We denote by C(B,©,e) the set of all IeC(B) with £ Xa<e for all Tg®. With these definitions, <zer Ptak's theorem [5] reads (2) The two following conditions on © are equivalent: a) There exists a strictly increasing sequence A2 czA2c= •• of finite subsets of A, and a sequence r„e© with An c Vn for all n; P) There exists an infinite subset B of A and an c>0 for which C(B,©,e) is empty. 00 Proof, a) Suppose that a) holds. Let B = \J An and IeC(B,©,e),£< 1. n= 1 Now and in what follows let N(I) denote the set of indices aeA for which Aa + 0. There exists an n0 with N(I)c=Ano. But then N(l)c=rno, so that Yj K= Z ^a=l> which contradicts the fact that IeC(B,©,e) aerV, aeA and that e<l. b) If A is a subset of A, let ©(A) denote the collection of all Te© with non-empty TnA. In particular let ©(a) denote the collection of all re© which contain a. We prove the following lemma: (3) Suppose that C(B,©,e) is empty and that B is infinite. If<P is a finite subset of A and if s1<e9 then there is a non-empty finite subset At of A which is disjoint from <$> and for which C(B ~ <t>, ©(A j), 82) is empty. It then follows from the definition of C(B, ©( A),e) that ©(AJ is non-empty. Proof. Suppose that C(B~<t>, ©(A^ej) is empty for given <t>, given s1<s and every finite non-empty subset AeA which is disjoint
328 § 24. Some results on compact and on convex sets from <P. We choose a finite non-empty subset Mj c B~<t>. Let I(1) be an element of the set C(B~<fc, ©(Mj), £2), which is non-empty, by hypothesis. We then form M2 = M1uN(I(1)). Then M2c=B~<t>, and we can choose an l(2) in C(B~<fc, ®(M2),ej, and so on. In this way we obtain a sequence \{n) = (?}"))eC(B^<P, ®(Mn), £2) with Mn = M1uN(I(1))u---uN(I("-1)). Suppose that re®. We form the sequence a„ = £/l(an):gl, and aeT assert that it contains at most one ocn^s1. Suppose that ap^a„ is the first of these a„. Since N(I(p)) c Mp+1, Mp+1nV is non-empty, so that re®(Mp+1). Now if q>p, re®(Mp + 1) c ®(Mq). Since i(^GC(B-0,®(Mq),81), a,<fi! for all q>p. Consequently if we form the mean — (l(1) + ••• + I(n)) for sufficiently n large n, then this lies in C(B~ <£,©,£), which contradicts the hypothesis of (3). c) Now suppose that /?) is satisfied. Then by (3) if st < £, C(B, ®(A 2), s^ is also empty. If we apply (3) again, then we obtain for £2<£i a non_ empty subset A2, disjoint from A x, for which C(B ^ A x, ®(A Jn ®(A2),£2) is empty. Repeating this procedure, we obtain a sequence A1?A2,... of pairwise disjoint non-empty sets for which the intersections ®(A1)n---n®(An) are non-empty. Now each ®(Afc) is the union of finitely many ®(afc), afeeAfc; there is therefore an al for which ®(a1)n®(A2)n---n®(AJ is non empty, for all m, and corresponding to it an a2 for which ®(a1)n®(a2)n®(A3)n---n ®(Am) is non-empty, for all m, and so on. Then ®(a1)n---n®(an) is non-empty, for all n. Now let A„= {(*!,..., a„}. The sequence A„ is properly increasing. For rne®(a1)n-"n®(an), however, A„ c r„, and so a) is established. The proof of Krein's theorem can be reduced to the case where E\X\ is ^-complete, as was shown in 5. By §20,6.(4), we can restrict ourselves to showing that C (M) is relatively weakly compact if M c E is. Using (1), Krein's theorem therefore follows from (4) // a bounded subset M of a real locally convex space has interchangeable double limits, then so does C (M). Proof. Suppose that the assertion is false. Then there is an absolutely convex weakly compact subset N of E and sequences xteC(M\ UjeN, with (5) lim lim u}xt — lim lim UjxA = c>0. i* J J i We denote by ji the finite value of sup \ux\. ueN, xeM There is a countable subset T of M with the property that all the x, lie in C (T). Suppose that u0 is weakly adherent to the sequence uy
6. Ptak's theorem 329 Using a diagonal procedure, we can choose a subsequence of the up denoted by u} once again, for which HmUjZ = u0z holds for all zeT. j We also have lim M7.x~w0xf for all i. j Since (5) holds, we can find a= ± 1 and an infinite set B of indices k for which (6) cmim ukxt — lim UqX-) ^fe for all /ceB. ^ i i ' For each zeT let l~(z) be the finite set of all indices; with (V) \(uj-u0)z\^^-. 4 Let © be the collection of all T(z), zeT. We now assert that c(b,©, — lis non-empty. Let us assume the contrary. Then by (2) there are index sets An = {/1,...,7n} and sets T(zn) with A„ c T(zn) for c "4 let y„ be a subsequence of the sequence z„ with the property that \im(ujq — u0)yn exists for each q. For the sequence y„eM and the n sequence vq = ^(ujq — u0)eN it now follows that \\mvqyn = Q for each rc, and that lim 1? yn\ ^ — for each q. By choosing a subsequence of the 1 " ' 8 I 1 e sequence 1; , we now obtain lim lim 1; )>„ ^ — and limlim i^yn==(). This contradicts the assumption that M has interchangeable double limits. Thus Cl B,©,— I is non-empty. Let I = (An) be a vector in this set, so that N(I)c=B. We set u = £ ^k(uk~uo)- By the definitions of / fi \ fceN(l) C B, ©, — and T(z), we then have, for each zeT, |wz|= Z ^(%-wo)zp Z 1 + 1 Z I 'fceN(I) ker(2) '/ceN (I)~r (z)1 — < — 4 ~ 2* all rc. By (7) it follows from this that \(ujq — u0)zn\ ^ — for n^iq. Now ^ 14 2/z+ I /J-< f(z) leN(I) 8 From this it follows that \uxt\ ^ — for all i.
330 § 25. Extreme points and extreme rays of convex sets Since N(I)c=B, we finally obtain the following contradiction from (6): — ^ lim \uxi\ = £ xJlim ukxt — lim w0x, 2 lN(I) \ i i )\ = [Yj Afc)<7(lim ukxt — lim u0xA ^fe. VN(1) J V »' * J For the results of this number I am indebted to a written communication from Professor Ptak, who has recently (cf. Ptak [6]) given a more extended account of his combinatorial method. § 25. Extreme points and extreme rays of convex sets 1. The Krein-Milman theorem. In the following, E may be either a real or a complex vector space. By (x,y\ however, we always mean the real open interval tx + (1— x)y, 0<t<1. Suppose that M is a subset of E. A point z of M is called an extreme point of M if it belongs to no open interval (x,y) cz M. If M is convex, this means that z never lies between two points of M. More generally, we clearly have: if z is an extreme point of the convex set M, and if z lies in the convex cover of the points xl9..., xn of M, then z coincides with one of the points xt. If M is absolutely convex and if z is an extreme point of M, then so is every point az with |c| = l. We denote the set of all extreme points of M by Ep(M). The concept of extreme point is a special case of a more general concept. A real linear manifold H c E is called a support manifold ofMif HnM is non-empty and if every open interval (x,j;)<=M which contains a point of if lies entirely in H, and thus lies in H nM. The zero-dimensional support manifolds of M are the extreme points; the support manifolds which are hyperplanes are the supporting hyperplanes of M (cf. § 17, 5.), provided that M is convex. It should be pointed out, however, that not every point of support (cf. § 17, 5.) need be a zero-dimensional support manifold, i. e. an extreme point. (1) Suppose that H is a support manifold of M in E. A real linear manifold H1 a H is a support manifold of M if and only if it is a support manifold of H nM. Proof. If H1 is a support manifold of M, it is also a support manifold of HnM, for H1r\(Hr\M) = H1nM is non-empty, and by hypothesis every open interval in HnM which meets Ht lies in H1nM = Hln(HnM).
1. The Krein-Milman theorem 331 Conversely, if Hx is a support manifold of HnM, HtnM is nonempty; further if (x,y) is an interval in M which meets Hx a H, then (x,y) is contained in H nM, since if is a support manifold of M. But then by the hypothesis on Hl9 (x,y) lies in H,. We now investigate the closed support manifolds of a compact subset of a locally convex space. (2) Suppose that M is a compact subset of a locally convex space. Further suppose that {Ha} is a system of closed support manifolds of M, which contains an Hpa JFfainJFfa2 whenever it contains Hai and Ha2. Then H = f]Ha is a closed support manifold of M. a Proof. The compact sets Ma = MnHa form a filter-base on M whose intersection H n M = f] Ma is non-empty, if is a closed support a manifold of M, for H is closed, and if (x,}/)cM has the point z in common with H then z belongs to all the Ha, so that (x,y)c=iJa, and consequently (x,y) czf]Ha = H. <x (3) Every closed support manifold H of a compact subset M of a locally convex space contains at least one extreme point of M. Proof. By (2), the closed support manifolds of M contained in H satisfy the hypotheses of Zorn's lemma, so that there is a minimal closed support manifold H0 of M contained in if. Suppose that H0 is not a point. We may suppose that oeii0 (this can always be obtained by a translation). By §20,7.(8) the compact subset H0nM of H0 has a closed supporting hyperplane H1 in H0. By (1), Hx is also a closed support manifold of M, which contradicts the minimality of H0. The Krein-Milman theorem now reads (4) Every compact convex subset M of a locally convex space is the closed convex cover of the set Ep(M) of its extreme points. We shall prove the rather more general statement (5) A compact subset M of a locally convex space has the same closed convex cover as the set Ep(M) of its extreme points. Proof. We assume that N1 = C{Ep(M)) is a proper subset of N=C(M). If z0eN~Nl9 then by §20,7.(1) there is a closed real hyperplane H which separates z0 and Nx strictly. There are points of M which belong to the same open half-space Rt of H as z0. By § 20, 7.(8), M has a closed supporting hyperplane H1 cz Rt parallel to H. By (3), ift contains an extreme point of M, and this cannot lie in Nl9 which is impossible. One half of the next theorem follows from (4):
332 § 25. Extreme points and extreme rays of convex sets (6) Suppose that K is a compact convex subset of a locally convex space. //McX, then Q(M) = K if and only if M => Ep(K). The other half results from Milman's theorem [2]: (7) Suppose that M is a set whose closed convex cover C (M) is compact. The extreme points of C (M) lie in M, and so they are also extreme points of M. Proof (following Bourbaki [6]). It is sufficient to show that every extreme point z of C (M) lies in each M+U, where U is a closed absolutely convex neighbourhood of o in the locally convex space. Now M is covered by finitely many sets xt+U, xteM, i=l,...,w. We set Mt=C (Mn(xf+[/)). Mi<= Xi+U9 and the Mh being closed subsets n of C (M), are compact, so that their convex cover C Mf is also com- n pact, by § 20, 6.(5). On the one hand, we now have C M{ c C (M), and on the other it follows from |J Mt => M that C Mf is a closed convex i=l i=l n n set containing M; consequently C M(- d C (M), so that C(M)= C Mf. i=l n i=l The extreme point z therefore has the form ^ afzf, zteMh Yjai=^ i= 1 at-^0. But this is only possible if z coincides with one of the points z{. Finally it follows from zeM{ and Mf <= xf + If that zeM+U. On going from M to C (M), extreme points can be lost, as the example of the set consisting of three collinear points shows. If M is compact and C (M) is not, then C (M) can have extreme points which do not lie in M. (For an example, cf. Bourbaki [6], Vol. 1, p. 85, Ex. 3.) The next result is a corollary of Milman's theorem: (8) Suppose that M is a set whose closed absolutely convex cover r~(M) is concept. Then every extreme point of I-(M) has the form az, |a| = 1, where z is an extreme point of M. Proof. Let M = [j ocM. M is compact, since it is the continuous image of the topological product of M and the set of all a with |a| = l. Since the circled cover of M lies in the convex cover of M, \~ (M) = C (M), by § 16,1.(2). By (7), every extreme point of \~(M) is therefore an extreme point of M, and a fortiori of some aM; it is therefore of the form az, where z is an extreme point of M. The set Ep(M) of a compact convex M is not in general closed. Suppose that in P3we are given a circle in the ^,?/-plane through the origin, together with the segment of the (-axis between (= — 1 and (= + 1, which meets it at o. The closed convex
2. Examples and applications 333 cover of this set has as extreme points the end-points of the segment and all points of the circle with the exception of o. Suppose that ^[^i] and £2[3^] are locally convex spaces, and that A is a continuous linear mapping from Ex into E2. The image Ax0 of an extreme point x0 of a set McEx need not be an extreme point of A(M). On the other hand, it follows easily from the definition of support manifold that the inverse image A{~l)(H) of a closed support manifold H of A(M) in A(El) is a closed support manifold of M. Applying (3), we obtain (9) Every extreme point of A(M\ where M is compact in ^[^J, is the image of an extreme point of M. 2. Examples and applications. The closed unit ball K of c0 has no extreme points: if x = (£n)ec0 and ||x|| = sup|£J = l, then we may re- n place some coordinate £k9 with |£fc|<l, alternately by £k + s and £k — e9 with 8 sufficiently small; x then lies between the two points of K obtained in this way. The closed unit ball K of/1 has the set of all <rzh i=l,2,..., |c| = l, as Ep(K). For if x=t=o-cf, we can divide the set of non-zero coordinates of x into two non-empty parts, to give two vectors xt and x2. x then 1 1 lies between s, and x9. 11*1 n iMi The unit ball of /°° has as extreme points all x = (£t) with |£f| = l for all i. If l<p<oo, then every boundary point of the unit ball of lp (respectively LP) is an extreme point. This follows easily from the remarks on Minkowski's inequality in § 14, 8. (respectively 10.). On the other hand, the unit ball of L1 has no extreme points: suppose b c that /eL1, \\f\dx=\\ we determine c in such a way that \\f\dx = ^. a a If we put j\ equal to If on \_a,c) and equal to 0 on [c,fc], and put f2 equal to 0 on [a,c) and equal to If on [c,fc], then / is the mid-point of the segment [/i,/2]> whose end-points both belong to K. The extreme points of the unit ball of L00 are the functions f(f) with 1/(01 = 1 almost everywhere. If R is a compact space and C(R) is the space of continuous real- (respectively complex-)valued functions on R (cf. § 14, 9.), then the extreme points of the unit ball K consist of all functions f(t) with \f(t)\ = 1. If, in addition, R is connected and C(R) is real, then it follows that /j = 1 and /2 = — 1 are the only two extreme points. We now investigate $Jl(R), the space of real measures on the compact space R.
334 § 25. Extreme points and extreme rays of convex sets By §24,5., the set R of point-measures Sx, xeR, is a subset of the weakly compact unit ball K of y)l(R) which is homeomorphic to R. A measure ue^Jl(R) is said to be positive if u(f)^0 whenever /^:0 on the whole of R. Let y)l + (R) be the set of positive measures, and let K* be the set of positive measures v with ||t;|| = 1. If veKi, t;(l)=l, so that Kf lies in the closed hyperplane H of 9K(R) given by m(1)=1. It follows from ||u||^l that m(1)^1, so that H is a supporting hyperplane of K. If ueKnH, ueKf, so that Kf=HnK. For if u(f)=-k<0 for ueKnH and for some /^:0, we could choose p>0 so that 1 — pf^O; we would then obtain u(l—pf)=l+pk>l, contradicting ||w|| = l. We now show that the points 5X are extreme points of Kf. We assume that there exist vl9v2eKi with (i) K=Uvi+v2). Suppose that g(x)eC(R) and that g(xo) = 0. We assert that v1(g) = v2(g) = 0. Since every such g is the difference of two non-negative functions vanishing at x0, it is sufficient to prove this for g^O. In this case, however, the assertion follows from v^g^O, v2(g)^0 and Sxo(g) = 0 = i(v1(g) + v2(g)). Now every feC(R) has the form f=f(x0)'l+g, with g(xo) = 0. But then v1(f) = f(x0)vl(l) + vl(g) = f(x0) = Sxo(f)9 so that v1=SX0 and v2 = SX0. The points 8X are therefore extreme points ofK^. The absolute polar of R in C(R) is the closed unit ball, and so K = \~(R), by the theorem of bipolars. By 1.(8), the only possible extreme points of K are the elements of R and of —R. The points of R are extreme points of Kf. Since Kf =KnH, they are also extreme points of X, by 1.(1). A corresponding argument applies to —R. Thus we have proved (2) Suppose that ^Jl(R) is the space of real measures on the compact space R. The extreme points of the unit ball K of 9W(K) are the point- measures Sx and their mirror-images —5X about o. The extreme points are shared between the two closed supporting hyperplanes u(\)=\ and m(1)=—1 of K. Remark. Similarly, the measures <rSx9 with |c| = l, are found to be the extreme points of the unit ball of the space of complex measures on R. The Banach-Stone theorem is now a simple consequence of (2): (3) Two compact spaces Rt and R2 are homeomorphic if and only if the real spaces C(RX) and C(R2) are norm-isomorphic.
2. Examples and applications 335 Proof. The condition is clearly necessary. Conversely, suppose that C(Rt) and C(R2) are norm-isomorphic. The dual spaces ^(jRJ and $ft(jR2) are then both weakly isomorphic and norm-isomorphic. In particular the sets of extreme points of the two unit balls are mapped onto each other in a weakly isomorphic way. By (2), every SyeR2 has as image either Sx or —Sx for some xeRl. It follows from this that the function 1 in QjRJ corresponds to a function cp in C(R2) with values + 1 and —1. Because cp is continuous, R2 divides into two compact open subsets Mx and M2, on which cp is equal to +1 and —1 respectively. The mapping which sends every f(x) in C(R2) to the function (p(x)f(x) is therefore a norm-isomorphism of C(R2\ under which cp is sent to the function 1. We may therefore suppose that 1 eC(Rl) is mapped into \eC(R2). R1 is then mapped in a weakly isomorphic way onto R2. By § 24, 5., Rt and R2 are therefore homeomorphic. The properties of the extreme points of the unit ball of 9W(jR) given in (2) characterise such spaces, as the following theorem of Arens and Kelley [1] (cf. Kadison [1], as well), which we give without proof, shows : (4) A real (B)-space E is norm-isomorphic to some C(R), with compact jR, if and only if the following hold: a) the extreme points of the closed unit ball K of the dual space lie in two supporting hyperplanes of the form u(a)= 1, u(b) = 1, where a and b are elements of E of norm 1 ; b) if the weak closure of a set of extreme points contains no two points symmetric about o, then it is contained in some supporting hyper- plane of K. For further characterisations of C(R), cf. Kadison [1] and Day [8]. The next result follows directly from the Krein-Milman theorem and the weak compactness of every closed equicontinuous set: (5) Every weakly closed convex equicontinuous subset of the dual E of a locally convex space E\%\ is the weakly closed convex cover of its extreme points. In particular the unit ball of the dual of a (B)-space has this property. From this it follows that (6) // E is a reflexive (B)-space, the closed unit ball of E is the closed convex cover of its extreme points. It follows immediately from (6) and the examples considered above that c0, L1 and the real space C(R\ R compact, infinite and connected, are not reflexive. Even more can be deduced from (5), however. (7) The spaces c0, L1 and the real space C(R\ R compact, connected and infinite, are not norm-isomorphic to a dual (B)-space.
336 § 25. Extreme points and extreme rays of convex sets For since their unit balls contain either no extreme points or only finitely many of them, the closed convex cover of the extreme points is at most finite-dimensional, for any Hausdorff topology, and so it cannot be the unit ball. The papers of Dixmier [1] and Ruston [6] deal with the question of when a (B)-space is norm-isomorphic to the dual of a (B)-space. Schatten [1] considers a further example, using the method of extreme points. 3. Variants of the Krein-Milman theorem. We begin with an example. Let M0 consist of a square with corners A,B,C,D, together with the coplanar semi-circular disc with diameter AB. A,B,C,D and the points of the open semi-circular arc R are the extreme points of M0. We can clearly leave out A and B, and still obtain the whole of M0 as the closed convex cover of C, D and jR. Quite generally, as has indeed been made explicit in 1.(6), Ep(K) can be replaced in the Krein-Milman theorem by any subset M of X, provided only that M =^ Ep(K). Variants of the Krein-Milman theorem are therefore obtained when the set Ep(K) is replaced by some such suitable set M. Suppose that K is a closed convex subset of a locally convex space E[Z]. If x0 is a point of support of K, i.e. a point through which at least one closed supporting hyperplane passes, let D(x0) denote the intersection of all the closed supporting hyperplanes through x0. D{x0) is a closed support manifold. We call the intersection S(x0) = D(x0)nK the supporting facet of K through x0. The supporting facet through a point is always a closed convex subset of K consisting only of points of support. If yeS(x0\ then S(y) a S(x0). A supporting facet of K is said to be minimal if it contains no proper subset which is also a supporting facet of K. If K is compact, every decreasing ordered set {^(xj}, with S(xa)aS(xp) if a>/?, has a non-empty intersection, which contains the supporting facet through each of its points. Thus by applying Zorn's lemma we obtain (1) If K is a compact convex subset of a locally convex space E[%~\, there is at least one minimal supporting facet in each closed supporting hyperplane of K. Two minimal supporting facets of K either coincide or are disjoint. We now obtain the Milman-Rutman theorem [1]: (2) Every compact convex subset K of a locally convex space E[%] is equal to C (M), where M is any subset which contains one arbitrary point from each minimal supporting facet of K.
4. The extreme rays of a cone 337 The proof proceeds in a completely analogous way to the proof of 1.(5), and will be left to the reader. Since every supporting facet of K is the intersection of a support manifold of K with K, by 1.(1) every extreme point of a minimal sup- porting facet is also an extreme point of K. Thus by (2) K=Q(M\ where M is any subset which contains one extreme point from each minimal supporting facet of K. In the example M0 above, the extreme points C, D and the points of the open semi-circle R are minimal supporting facets, while the supporting facet through A is the segment AD; it is therefore quite possible for the supporting facet through an extreme point to fail to be minimal. A minimal supporting facet need by no means consist of a single point. If we add another semicircular disc, with diameter CD, to the set M0, we obtain another compact convex set Mi for which the two segments AD and BC are minimal supporting facets. An essentially stronger result has been obtained by Klee [9] for normed spaces. Following Straszewicz [1], a point x0 of a subset M of a locally convex space E [%] is called an exposed point of M if x0 is a point of support of a closed supporting hyperplane H of M, for which HnM = {x0}. Every exposed point of M is thus an extreme point, but not conversely, as for example the point A of the set M0 shows. In particular, exposed points of a set are zero-dimensional supporting facets. The following theorem of Klee's [9] is thus an extension both of 1.(4) and of (2), for normed spaces. (3) Suppose that K is a compact convex subset of a normed space. The set of exposed points of'K is dense in Ep(K), so that K is the closed convex cover of the set of exposed points of K. Reference must be made to Klee [9] for the proof, and for a collection of further results about exposed points. 4. The extreme rays of a cone. An analogue of the Krein-Milman theorem can be proved for cones, once a further concept has been introduced. Suppose that K is a convex subset of a vector space E. Let h be a real open half-line contained in K; it is called an extreme ray of K if every open interval in K which meets h lies wholly in h. It follows from this that the straight line g which h defines meets K in a set which consists only of h and possibly the end point a of h. If a also belongs to X, then a is an extreme point of K. If C is a convex cone in E with vertex x0, and if x is a point of the cone other than x0, then x lies on one of the generators of the cone, namely the open half-line from x0 through x. It follows directly from this that x0 is the only possible extreme point of C, and that x0 is an extreme point of C if and only if C is proper and pointed. The question of which generators of C are extreme rays can be answered in the following way: 22 Kothe, Topological Vector Spaces I
338 § 25. Extreme points and extreme rays of convex sets (1) Suppose that C is a proper convex cone with vertex x0. If a generator h is an extreme ray, and if a hyperplane H cuts the generator h in one point y =Nx0, then y is an extreme point of H nC. Conversely if there is a hyperplane H which cuts the generator h in an extreme point of H nC, then h is an extreme ray of C. Corollary. // M is a convex set contained in a hyperplane H, and if C is the cone generated by M with vertex x0 outside H, then the generators through the extreme points of M are precisely the extreme rays of C. Proof. The first part of (1) follows directly from the definition of extreme point and extreme ray. Conversely suppose that h meets H in an extreme point y0 of HnC. Suppose that h is not an extreme ray. Then there is an open interval (x1,x2) cz C which meets h but which does not lie in h. It is easy to see that we can suppose that the point of intersection is y0. But then the open planar surface bounded by the generators x0x1 and x0x2 is contained in C, contains h9 and meets H in an open interval containing y0; y0 is therefore not an extreme point of H n C, contrary to hypothesis. We now consider cones in a locally convex space £[£]• (2) A proper pointed convex cone C in E[%] is locally compact if and only if it is the pointed cone generated by a compact convex set M which does not contain the vertex. M can be chosen to be a subset of a closed hyperplane. Such a cone is always closed. In the proof we can suppose that the vertex is the point o. a) Suppose that C is proper, pointed, convex and locally compact. Then there is a closed convex neighbourhood U of o in £[T] for which U n C is convex and compact. Let M be the intersection of the boundary of U with C, and let K = Q{M). From M^UnC it follows that K ci U n C, so that K is convex and compact. If yeC, j/4=o, then the generator through y contains a boundary point of V, and so a point of M; C is therefore the cone generated by M, and a fortiori is the cone generated by K. We still have to show that o does not belong to K. If o belonged to K, o would be an extreme point of K, since C is assumed to be proper. By Milman's theorem (1.(7)), o would then have to be an extreme point of M, and this is not possible. b) Suppose that C is the pointed cone generated by the convex compact set K with vertex o$X. By §20,7.(1), there is a real closed hyperplane H which separates o and K strictly. H intersects every generator of C in a point which is different from o. Let K = H nC. K is a subset of the set C1 of all py, pe[0,1], yeK. This set, being the continuous image of the compact set [0, l]xX, is also compact. Thus K = H nCl is convex and compact, and C is the cone generated by K.
5. Locally compact convex sets 339 c) Let C be the cone generated by K = HnC, as in b), and let z=f=o be a closure point of C. The ray from o through z meets H, for otherwise z would lie in the hyperplane H0 through o parallel to H, and z would therefore be a closure point of the part C1 (the set of all py, pe[0,1], yeK) of C which lies on the same side of H; but since Ci is compact, JL/0nC1=o. Since therefore the ray from o through z meets H, there is a hyperplane Hi, parallel to H, such that z is an interior point of the closed half-space Rl defined by H1 and containing o. z must be a closure point of the part of C lying in jRl5 and this has the form Q (the set of all ay, <7e[0,/], yeK) and is therefore compact. Hence zeC, and so every proper pointed convex locally compact cone is closed, by a) and b). d) Finally suppose that C is the cone generated by a compact convex set M which does not contain o. By b), we can replace M by a set K lying in a closed hyperplane H. Every point of C has the form py, p^O, yeK. If we construct the hyperplane H2 parallel to H through (p + \)y, py is an interior point of the closed half-space R2 defined by H2 and containing o, and so R2nC is a compact neighbourhood of py in C. Thus C is locally compact. We now have the following analogue of the Krein-Milman theorem. (3) Every proper pointed convex locally compact cone C in a locally convex space is the closed convex cover C {Es(C)) of the set Es(C) of its extreme rays. Proof. By (2), C can be represented as the pointed cone with vertex o generated by a compact convex set K lying in a hyperplane H. By the corollary of (1), the extreme points of K lie on extreme rays of C; by the Krein-Milman theorem, K is contained in C (Es(C)). Any hyperplane parallel to H which cuts C other than in o can be used to produce C, and so every non-zero point of C lies in C (Es(Cj); from this it follows that C cz Q (Es(C)). C is closed, by (2), and so we also have the converse inclusion C (Es(Cj) cz C. The set Wl + (R) of positive measures on the compact space R forms a proper pointed convex weakly closed cone in Wl(R). Wl + (R) is the cone generated by Kj", the weakly compact convex subset of positive measures of norm 1. We have shown that the point-measures Sx are extreme points of K±. Since K± =KnH, and H is a closed support hyperplane of the unit ball K, Kf has no further extreme points, by 1.(1) and 3.(2). The extreme rays of 5R + (jR) are thus the generators defined by the points dx, and 9Jt + (K) is the weakly closed convex cover of these extreme rays. 5. Locally compact convex sets. The Krein-Milman theorem and its analogue for cones, which we have just obtained, are special cases of a more general theorem of Klee's [8]: 22*
340 § 25. Extreme points and extreme rays of convex sets (1) Every closed convex locally compact subset K of a locally convex space which contains no straight line is the closed convex cover of the set of its extreme points and extreme rays: K = C {Ep(K)u Es(K)). Before proving this, we establish two lemmas: (2) // K is closed, convex and locally compact, K is compact if and only if it contains no half-line. We need only show that if K is not compact then it contains a half- line. We shall prove rather more, namely that through each point of K there is a half-line contained in K. We can take this point to be o. There exists a closed convex neighbourhood U of o for which U n K is compact. The intersection B^ of the boundary of U with K is also compact, and it does not contain o. Let Bn be the intersection of n U with K, multiplied by \/n. Clearly Bl ^ B2 => •••. We shall show that no Bk is empty. If Bk were empty, we would have K = (kU)nK. It would then follow from (kJJ)nK <=/c((7r\K) and the compactness of [/nK, and thus of k(UnK), that K would be relatively compact; since K is closed, K would therefore be compact, contrary to hypothesis. The sequence J3X => J32 => — of compact sets Bt therefore has a nonempty intersection. If z=t=o is an element of this intersection, every nz belongs to K, and so the ray from o through z lies in K. (3) If K is closed, convex and locally compact, and if K contains no straight line, then K has at least one extreme point. We may assume that K is not compact. By (2) there is at least one point x0 from which there is a half-line contained in K. The collection of all half-lines from x0 which are contained in K forms a proper, closed, convex cone C. This cone is also locally compact, since each point of C has a compact neighbourhood in K, and therefore also in KnC. By 4.(2) there is a closed hyperplane H^ which has only the point x0 in common with C. Let jRt be the closed half-space defined by Hi which does not contain C. Rx has a locally compact intersection K1 with K. If Kx were not compact, then there would be a half-line through x0 lying in Kl9 by the proof (2). But this contradicts the definition of C. If the compact convex subset Kx of K lies in Hl9 then every extreme point of Kx is also an extreme point of K. If K1 <£ H{ then K± has an extreme point which does not lie in Hu and which is therefore an extreme point of K. Proof of (1). We assume that A= C (Ep(K)uEs(K)) is a proper subset of K. Then by §20,7.(1) (applied to a point of K~A, and A), there is a closed hyperplane H which meets K, but not A. KnH satisfies the hypotheses of (3), and so it possesses an extreme point x0, which
5. Locally compact convex sets 341 is not an extreme point of K, since it does not lie in A. There is therefore a straight line g, cutting H in x0, whose intersection D with K contains x0 as an interior point. D is either a segment or a half-line. Suppose first that D is a segment [x1?x2]. We assert that xt and x2 are extreme points of K. If xt were not an extreme point, there would be two points yl9y2eK9 not lying on g, with x1e(yl9y2). If z is a point of D on the side of H which does not contain xl9 then the triangle z, yl9y2, which lies in K, would meet H in an interval containing x0 as an interior point, and this is not possible, since x0 is an extreme point of KnH. Since x0 lies between the two extreme points x1 and x2 of K9 x0 must belong to A; this gives a contradiction. If D is a half-line, then the same argument shows that every open interval in K which contains a point of D must lie in D. But then D is an extreme ray of K. Since x0 lies in D9 x0 must again belong to A, which is impossible. Thus (1) is proved. Suppose that M is a subset of a locally convex space E\%~\ and that h is a half-line x0 + py9 p ^0, in E. The half-line H is called an asymptote of M if for each p>0 and each ^-neighbourhood U of x0 + py there exists a zeM for which [x0,z] n (7 is non-empty. Suppose that K is convex, locally compact and closed, and suppose that K contains no straight line. Suppose further that M is a subset of K with M =5 Ep(K)9 and suppose that every extreme ray of K is an asymptote of M. Since the end point of an extreme ray of K is an extreme point of K9 and so lies in M, it follows easily that C (M) contains every extreme ray of K as well. Thus C (M) is equal to K9 by (1). This establishes one half of the following generalization of 1.(6): (4) Suppose that K is a closed convex locally compact subset of a locally convex space, which contains no straight line. A subset M of K satisfies the relation C (M) = K if and only if M => Ep(K) and all the extreme rays of K are asymptotes of M. Reference must be made to Klee [8] for the proof of the other half, which is a not completely straightforward generalisation of Milman's theorem. In 3. we introduced the concept of an exposed point. An exposed ray H of a closed convex subset K of a locally convex space is a closed half-line which is the intersection with K of a closed supporting hyper- plane H of K. The corresponding open half-line is an extreme ray of K. Klee [9] has proved the following stronger form of (1), corresponding to 3.(3): (5) Suppose that K is a closed, convex, locally compact subset of a normed space which contains no straight line. K is the closed convex cover of the set of exposed points and exposed rays of K.
342 § 26. Metric properties of normed spaces § 26. Metric properties of normed spaces 1. Strict convexity. Although every normed space is a topological vector space, it has an even richer structure. There is a collection of properties which are defined in terms of the metric, with the result that they are preserved under norm-isomorphism, but not under topological isomorphism. These properties are important above all for various applications in analysis. In this paragraph we shall investigate some of the concepts of this kind. First suppose that £[£] is locally convex and that K is a closed convex £-body with o as interior point, and with boundary S. K is said to be strictly convex, or rotund, if every boundary point of K is an extreme point. A normed space, and its norm, are said to be strictly convex if the closed unit ball is strictly convex. (1) // K is a closed convex X-body, the following conditions are equivalent : a) K is strictly convex; b) S contains no line segment ; c) every supporting hyperplane meets K in at most one point; d) distinct boundary points have distinct supporting hyperplanes; e) every boundary point of K is an exposed point. If K is the closed unit ball of a normed space, we can also add 0 if M = |M| = 1 and x*y, then ||i(x + y)||<l; g) if \\x + y\\ = \\x\\ + \\y\\ and y+o, then x = oiy for some non- negative a. Proof, b) follows from a). Suppose that b) is satisfied, but not a). Then there is a zeS and a segment [x,y] c K with z as interior point. If x were an interior point of X, z would be an interior point of the convex cover of y and a neighbourhood U <=K of x, and this is contained in K. Thus [x,y]<=S, contradicting b). c) is necessary; for if a hyperplane of support H met K in more than one point, HnX c S would contain a line segment, contradicting b). Conversely if z is a boundary point of K, z is point of support of a closed hyperplane H (cf. § 17, 5.(1)) and it follows from c) that z is a minimal support manifold, and so is an extreme point. d) follows from c). b) follows from d), since there is a supporting hyperplane through every segment in S (§ 17, 2.(2)). Since there is a supporting hyperplane through each point of S, c) and e) are equivalent. If K is the closed unit ball of a normed space, and if f) holds, then b) follows. On the other hand, f) follows directly from a).
2. Shortest distance 343 Suppose that g) is satisfied. Ifx, yeE and if ||x|| = \\y\\ = Ili(x + j;)|| = l, then ||x + j;|| = ||x|| + ||j/||; it follows from g) that x = ay, and we must then have x = y. f) therefore follows from g). We now assume that f) holds. If ||x + j/|| = ||x|| + \\y\\ for x, yeE, and if ||j/||^||x|| and x,y4=o, then 1 x y \ \\x\\ \\y\\ > \ x y \M\ \\x\\ — y IMI ' y 1 " IMI1 -w(si-m)=2- It now follows from f) that x/||x|| = y/\\y\\, i.e. g) is satisfied. For the examples discussed in § 25, 2. we have the following results: the spaces lp and LP are strictly convex for 1 <p<oo, while the spaces c0, l\ /°°, L1, L00, C(K) and Wl(K) are not strictly convex. 2. Shortest distance. Suppose that M is a subset of the normed space £, and that x0 is a point of E. The quantity |x0,M| = inf ||x0 — y\\ yeM is the distance of x0 from M. If the infimum is attained for some y0eM, then we call this point a nearest point to x0 in M. We shall be concerned with the question of when a nearest point exists, and when it is unique. (1) Suppose that E is a normed space, and that M is a closed convex locally compact subset of E. Then there is always at least one nearest point to x0 in M. Proof. The closed unit ball K of E and M are weakly closed. For sufficiently large p^O (x0 + pK)nM is non-empty. This set is weakly closed and contained in M, so that it is weakly locally compact. By § 25, 5.(2) it is therefore weakly compact. The intersection of all the nonempty (x0 + pK)nM is thus non-empty, and is equal to (x0 + p0K)nM, where p0 is the least of these p. All the points of (x0 + p0K)nM (and no others) are nearest points to x0 in M. (2) If E is a reflexive (B)-space and if M is a closed convex subset of E, then there exists at least one nearest point to x0 in M. This follows in the same way as (1), for the closed bounded sets (x0 + pK)nM are all weakly compact. The next result shows the significance of strict convexity for this problem: (3) // E is strictly convex, and if M is a convex subset of £, then each x0 has at most one nearest point in M.
344 § 26. Metric properties of normed spaces If there were two nearest points yt, y2 at distance p0 from x0, then \(yl +y2)£M, being the midpoint of two boundary points of the strictly convex set x0 + p0K, would, by 1.(1)f, have distance less than pQ from x0, and this is not possible. Clearly strict convexity is also necessary for (3) to hold. If E is strictly convex and if M satisfies the hypotheses of (1) we can define a mapping from E onto M, the nearest-point mapping PM(x), which sends each xeE to its uniquely determined nearest point in M. (4) // E is strictly convex and if M is a closed convex locally compact subset of E, then the nearest-point mapping PM{x) is continuous. We must show that PmW-^m^o) always follows from x„->x0. If \\xn-x0\\^s, then dn= inf ||x„-.y|| ^ inf||x0-y|| + ||x0-xj = d0 + e. yeM From this it follows that \\x0-PM(xn)\\ ^ \\xn-PM(xn)\\ + ||x0-xj|^ + e^rf0 + 2e. Thus PM(xn) lies in C(e) = (x0 + (d0 + 2s)K)nM. This set is compact. The diameters of these sets C(e) must tend to 0 as e-^O, for otherwise there would be a nearest point to x0 in M different from Pm(xo)- The PM{xn)eC(e) therefore converge, as e-^O, to PM(x0) = n cm. £ >0 In particular, if M is a finite-dimensional linear subspace of a normed space E there is always at least one best approximation of xeE by elements of M, by (1). If, further, M is strictly convex, there is only one such best approximation, and it depends continuously on x. If M is a closed linear subspace of /2, the nearest-point mapping is the orthogonal projection on M, and so it is linear. In the general case the nearest-point mapping need not be linear, even when M is a closed linear subspace. We now investigate the behaviour of |x,M|=/(x) with regard to the weak topology. We first make two general remarks about normed spaces. Suppose that £ is a complex normed space, and that Er is the same space considered as a real normed space. If ueE\ then $iue(Er)\ and we assert that (5) IMI = I|5R"I|. Once again, let K be the closed unit ball of E and of Er. If ux = rei(p, then \ux\ = ($iu)(e~i(px). It follows from this that \\u\\ = sup \ux\ ^ sup \(9lu)y\ = ||5Rm||, xeK yeK and since ||9tw||^||t/||, (5) follows. We now prove a generalization of § 17, 6.(2):
3. Points of smoothness 345 (6) Suppose that M is a closed convex subset of the normed space E, and that x0 has distance \x0,M\ = d from M. Then there is a uQeE' with ||w0|| = l and 9l(u0{x0 — y))^.d for all yeM. Because of (5) and § 16, 3.(1), it is sufficient to prove this for a real normed space. Further, we can take x0=o. We denote the open ball ||x|| <p in E by Klp. By hypothesis the two convex sets Kld and M have no point in common. By § 17,1.(4) they are separated by a closed hyperplane given by u0x = p,p>0, with ||w0|| = l. In particular sup u0x = \\u0\\d^p, so that p^d. But then u0y^d for all yeM. xeK'" (7) If M is a closed convex subset of a normed space E, f(x)= \x,M\ is a weakly lower semi-continuous function on E. Proof. Suppose that u0 is an element of E which corresponds to x0, as in (6). Then f(x0) = d= inf||x0-.y||^ inf ^(u0(x0-y))^d, yeM yeM so that /(x0) = inf5R(M0(x0 —y)). Now if x is in the weak neighbourhood \u0(x0 — y)\<c, we have /W=|x,M|^inf^(Wo(x-3;))^inf^(Wo(x0-3;))-|^K(x-x0))| ^/(x0)-c, so that f(x) is weakly lower semi-continuous at x0. The distance between two subsets Ml and M2 of E is defined by inf ||x — j;|| = |M1,M2|. From (7), we obtain xeMi, yeM2 (8) Suppose that Mr is a weakly compact subset, and M2 a closed, convex, weakly locally compact subset of the normed space E. Then there are two points xleMl and x2eM2 with \\xt — x2\\ = \Ml,M2\. Proof. The function |x,M2|, which is weakly lower semi-continuous, by (7), takes its minimum value |Ml5M2| at a point xr of Ml9 by §6, 2.(6). By (1) there is an x2eM2 with ||x1-x2|| = |M1,M2|. Remark. If £ is a reflexive (B)-space, Mx can be taken to be an arbitrary bounded closed set, and M2 to be an arbitrary closed convex set. 3. Points of smoothness. We obtain a concept which corresponds dually to the concept of extreme point, by making the following definition. A boundary point x0 of a convex subset K of a vector space E is called an algebraic point of smoothness of K if at most one supporting hyperplane passes through x0. If £p] is locally convex, and if at most one closed supporting hyperplane passes through x0, then x0 is called a point ofsmoothness of K.
346 § 26. Metric properties of normed spaces We again consider a closed convex 2-body Kbo in £[2] with boundary S; let K° be the convex weakly compact subset of E polar to K. Every supporting hyperplane of x0eS is closed, and has the form y{{ux)=l, with ueK°. By § 17, 5.(1), there is a closed supporting hyperplane through each point x0 of S. If x0 is a point of smoothness of K, there is therefore exactly one supporting hyperplane through x0. In this case we speak of a tangent hyperplane. The points of smoothness of S are therefore the points at which a tangent hyperplane exists. If to each point of S we make correspond all those points ueK° which define supporting hyperplanes (iR(ux)=\ through the point, we obtain a many-valued function from S into K°. For this we have (1) a) To every point of smoothness of S there corresponds just one point of K°, which is extreme; b) To every point S which is not a point of smoothness there corresponds at least one non-extreme point; cj to every point of S which is not an extreme point there corresponds at least one point of K° which is not a point of smoothness. Proof, a) If the point u0 corresponding to the point of smoothness x0 were not an extreme point, then there would be an interval [wl5w2] <= K° which contains u0 as an interior point. But then $i(u0x0)= 1 would be a point of \_(iR(ulx0), 9i{u2x0)]. It would then follow from <^R(u1x0)S\ and 9?(w2x0)=l tnat (R(u1x0) = <R(u2xQ)= 1, so that (iR(ulx)=\ and y{(u2x)=\ would be two further supporting hyperplanes through x0, and this is not possible. b) If x0 is not a point of smoothness, let (R(ulx)=\ and ^R(u2x)=\ be two distinct supporting hyperplanes through x0. Then [wl9w2] is contained in K°, and for each interior point u0 of this line segment we also have $R(w0x)=l, i.e. u0 defines a supporting hyperplane through x0; uQ is not an extreme point of K°. c) Suppose that x0 is an interior point of [xl5x2] <= S. There is a supporting hyperplane (iR(u0x)=\ containing [xl9x2]. But then all supporting hyperplanes of K° of the norm (R(ux)=l with xe\_xl,x2] pass through u0, so that u0 is not an algebraic point of smoothness of K°; u0 is also not a £s(£)-point of smoothness of K°. A convex 2-body is said to be smooth if all the points of its boundary are points of smoothness. In particular a normed space, and its norm, are said to be smooth, if the closed unit ball is smooth. From (1) b) and c) we obtain immediately that (2) // the strong dual E of a normed space is strictly convex (respectively smooth), then E itself is smooth (respectively strictly convex). Complete duality clearly holds in the reflexive case:
4. Weak differentiability of the norm 347 (3) A reflexive (B)-space is strictly convex (respectively smooth) if and only if its strong dual is smooth f respectively strictly convex). Day [6] has given an example of a non-reflexive strictly convex (B)-space whose dual is not smooth (cf. 9.). (4) Linear subspaces of strictly convex (respectively smooth) normed spaces are again strictly convex (respectively smooth). This is clear for strict convexity. If H is a subspace of E and if y0 is a boundary point of the unit ball of H, then it follows from the Hahn- Banach theorem that if y0 is a point smoothness of the unit ball of E then y0 is also a point of smoothness of the unit ball of H. The second part of (4) follows from this. (5) If E is a reflexive strictly convex ( respectively smooth) (B)-space, then every quotient space E/H is also strictly convex (respectively smooth). By §22, 3.(lb), (E/H)' is norm-isomorphic to a closed linear sub- space of E\ which is smooth (respectively strictly convex), by (3) and (4). The assertion now follows from (2). (5) does not hold for quotients of arbitrary normed spaces: cf. Klee [10]. It follows directly from (2) and the examples in 1. that none of the spaces li = (c0)', /00 = (/1)/, L°=(Ll)f, 9W(JC) = C(JQ' is smooth. Further examples follow in number 5. 4. Weak differentiability of the norm. The question of when a boundary point of the unit ball of a normed space has a tangent hyperplane is connected with the differentiability of the norm at this point (cf. Mazur [2], [3]). Let q(x) be the Minkowski functional of a convex 3>body Cbo in the locally convex space £[£]. q(x) is continuous and has properties (a), 0S)and(y)of§16,4.(2). q(x0 + ty) — q(x0) Let A(x0,j;,t) = be the difference quotient of q(x) at x0 in the direction j;. (1) A(x0,j;,t) is monotonic increasing for t>0. Proof. If 0<tt <t2, we have t2q(x0 + tly) = q{t2x0 + t2tly)^tlq(x0 + t2y) + (t2-t1)q(x0), so that h(q(x0 + tly)-q(x0))^t1(q{x0 + t2y)-q(x0)). (2) h(x0,y,t) is bounded below, for t>0, and we have h(x0,y,t) ^-q(-y).
348 §26. Metric properties of normed spaces This follows directly from q{x0) = q(x0 + ty-ty)^q(x0 + ty) + tq(-y). The existence of the right derivative q'+(x0,y) = lim h(x0,y,t) of q(x0 + ty) at t = 0 now follows from (1) and (2). From (2) we obtain (3) q'+(x0,y)^-q(-y). It follows from A(x0,j;, —t) = — A(x0, — y,t) that A(x0,j;,t) is also monotonic increasing and bounded above for negative t; for the left derivative q'-(x0,y) = lim A(x0,j;, — t) we have (4) q'-(xo,y)=-q'+(x0,-y)- It follows from 2q(x0)^q(x0 + ty) + q(x0 — ty) that A(x0,j/, — t) ^A(x0,j/,£), and, as £->0+, this gives g'-(x0,.y)^tf+(*()>.)>)• From this we obtain (5) A(x,j;,£) is monotonic increasing for all t, and we have the inequality ,,, q(x0-ty)-q(x0) .,q{x0 + ty)-q{x0) (6) — ^ ^(Xo^tf+Ow) ^ , r>0. The next proposition gives further properties of q'+. (7) For fixed x0, q'+(x0,y) is a continuous positive homogeneous subadditive function on E[_%~]. Proof. It follows directly by taking limits in A(x0,<rj;,t) = (j£L{x0,y,(Tt), for <r>0, that q'+(xQ,(jy) = (jq'+(x0,y). By subtracting 2q(x0) from the inequality 2qU0 + -(yi+y2)\ ^ q(xQ + tyl) + q(xQ + ty2) and dividing by t, we obtain AUo^+j^-j ^ A(xo,3;1,0 + A(xo,3;2,r). As f->0, this gives q'+(x0,yl+y2)^q'+(x0,yl) + q'+(x0,y2). Finally the inequality (8) q'+(xQ9y)^qiy) follows from A(x0,j;,t)^ql — J + q(y) — q[ — ) = q(y\ and this, to-
4. Weak differentiability of the norm 349 gether with (3), establishes the continuity of q'+{x0,y) at o; the ideas of the proof of § 16,4.(7) then give continuity for all y. We observe that at x0 itself we clearly have q'+(xQ,xQ) = q'_(xQ, x0) = l. If the two derivatives q'-(x0,y) and q'+(x0,y) coincide, the common value q'(x0,y) = limA(x0,y,t) is called the weak or Gateaux deriv- ative of q at the point x0 in the direction j;. In particular q(x) is said to be weakly differentiable at x0 if q'(x0,y) exists for each yeE. For this it is necessary and sufficient, by (5), for ,m r q(x0 + ty) + q(x0-ty)-2q{x0) (9 hm = 0 t-o+ t to hold for all yeE. (10) // q(x) is weakly differentiable at x0, the function defined by uy = q'(x0,y) is a real continuous linear functional on £[£]. The linearity and continuity of u follows without difficulty from q'-(x0,y) = q'+(x09y), (4) and (7). The next theorem gives the connection with supporting hyperplanes: (11) Suppose that Cbo is a closed convex %-body in the locally convex space £p], with Minkowski functional q(x). A boundary point x0 of C is a point of smoothness if and only if q(x) is weakly differentiable at x0. The tangent hyperplane through x0 is then given by q'(x0,y)=\. More generally we have: if ux=\ is a real supporting hyperplane through x0, u satisfies the inequality 02) -<?'+(*()> -y)S-uy^q'+(x0,y) for all yeE. Conversely if yQeE and if —q'+(xQ,—yQ)^y^q'+(x0,y0) then there is a real supporting hyperplane vx=\ through x0 for which »yo = y- It is sufficient to prove the second part of the assertion. Suppose that ux=\ is a supporting hyperplane through x0. Then ux^q(x), and it follows from 1 +tuy = u(x0 + ty)Sq(x0 + ty), t>0, *u . ^<l(xo + ty)-q(xQ) that uy ^ , so that uy^q + (xQ,y). It then follows from w( — y)^q'+(x0, — y) that —u( — y) = u(y)^—q'+(x09—y). Thus (12) is established.
350 § 26. Metric properties of normed spaces We now define a linear functional on the real linear subspace H of E spanned by x0 and y0 by setting l(z) = l(oLx0 + Py0) = oL + Py. For sufficiently small />0 we have \ 1+OLt J J from wjiich it follows that q'+(x0,z) = (x + q'+(x0,Py0). From the definition of y, we have PySq'+(x0,Py0), and we therefore obtain l(z) = ot + Py^ot + q'+(x0Jy0) = q,+ (x0,z). It now follows from (8) and the Hahn-Banach theorem that there exists a real linear functional v with vy^q'+(x0,y)^q(y) for all yeE. As a special case of (11) we have (13) A normed space is smooth if and only if its norm is weakly differ- entiable at every point except the origin. Since weak differentiation always takes place in a two-dimensional subspace, we have (14) A normed space is smooth if and only if every two dimensional linear subspace is smooth, and if and only if the norm is weakly differen- tiable (except at the origin) in each of these subspaces. 5. Examples. 1. We now determine all the points of smoothness of the unit ball of C(R), R compact (cf. Banach [3], p. 169). In § 25,2. we determined the extreme points of the unit ball of Wl(R). These are the points adx, with x arbitrary in R and |<r| = l. By 3.(l)a) the supporting hyperplane through a point of smoothness f0 must have the form 3i((rdx(f))=l. But there is just one hyperplane of the form 3l(<jSx(f))=l through a point f0, with ||/0II = 1> if anc* only if the function f0 takes a value of modulus 1 at just one point x0. The norm p(f) = sup|/(t)| is therefore differentiate only at these teR points f0, and we have (i) p'(fo,g) = m7J^)K(g))- 2. A similar idea can be used for L1. The extreme points of the unit ball of L00 are the functions h(t) with |/i(t)| = l almost everywhere. If now J|/(r)|dr=l, f(t) can only lie on the hyperplane 9l^h(t)f(t)dt)= 1 fit) if f(t) is non-zero almost everywhere and if h(t) = ——- holds almost \ J wl everywhere. From this it follows that the points of smoothness of the unit ball of L1 are those functions f0(t) with ||/0II = 1 an<^ /oW + 0 q(x0 + tz) — q(x0) 1+at = a H t t
5. Examples 351 almost everywhere, and that at these points f0 the weak derivative of the norm has the form /oW (2) "^-rm^r This argument assumes that it has already been shown that {I})' = L? (cf. 7.). A proof can be given without making this assumption by using an argument similar to that of the next example. 3. We now consider IF, 1</?<oo, (cf. Mazur [3]). We calculate the derivative of the norm directly. ^-[(jl/oW + ^WI^t)1/P]h = o=--(il/olP^r*-^(Il/o + ^lP^=o, an p ah whenever the derivative on the right hand side exists. The difference quotients of the convex function \f0 + hg\p are monotonic increasing with h; by Lebesgue's theorem, we can therefore take limits under the integral sign as /i-»—0 and /i->+0; a derivative therefore exists as /i-»0, provided that \f0 + hg\p is differentiable at h = 0. If we put \f0 + hg\p = [(f0 + hg)(f0 + hg)y<\ we obtain dh\f0+hg\%=o=p\for2mo9)=p\for1*WjLlg Combining these results we see that the weak derivative of \\f\\p = p(f) = ($\f\pdt)1/p is equal to (3) p,(/o^) = ll/ollp' i/»r'*(^K and it therefore exists at every point except the origin. Consequently every boundary point of the unit ball is a point of smoothness, and (3) gives its supporting hyperplane. This result can be proved much more simply if we use the fact, not 1 1 proved until 7., that (LP)' = I3, where — + — = 1. Since 13 is strictly P <1 convex, Lp is smooth, by 3.(3). If f0eLp and ||/0llp=1 then tt^I/o f° IfoK1 f \ \J0\ Jo is in 13 and has L^-norm 1, so that \fQ\p * 911 g \dt= 1 is the supporting hyperplane through f0. J ^Jo 4. The points of smoothness of Z1 can again be found using the extreme points of/00. From the results of § 25, 2. we find that the ^-points
352 § 26. Metric properties of normed spaces of smoothness are just those points * = (£„) for which all the coordinates £n are non-zero. At such a point the derivative of the norm is given by (4) p'(*,i)) = SRf I enrin), where fi« = 77^- 5. In a similar way it follows that /p, 1 <p< oo, is smooth, and that the derivative of the norm is given by oo F (5) p'(s,D) = K~P £ l&,r'MM,), where c„ = -|l n= 1 I Cm I 6. We shall show that no boundary point of the unit ball in L00 is a point of smoothness. We can restrict our attention to L°(I), with / = [0,1], as the case / = ( — 00,00) can easily be reduced to it. If ||/0||=esssup|/0(/)| = 1, there is at least one cr, with |c| = i, for which \<j — f0(t)\^e holds on a set of positive measure, for each e>0. There is then a sequence /„ of pairwise disjoint measurable subsets of /, each of positive measure and with \a— f0(t)\^- for all tel„. d f ' n The functions un(f) = f(t)dt are continuous linear functional Mn) J In on La with |1— un(f0)\ ^ — and ||uj = l. Each of the two sequences n u2n-i and u2n has a weakly adherent point u{{) (respectively u{2)) in the weakly compact unit ball of (L00)'. It follows from u{1)(f0) = ui2)(f0)=l that they define two supporting hyperplanes through f0. These are distinct: if %n is the characteristic function of In then u2k[ Y, %2n) = ^ so that m(2)(Z^2«) = ^ while w2fc-i(I>2,,) = 0, so that w(1)(I%2m) = 0. 7. The space /°°. Suppose that ate/00 and that ||x|| = l. If there is a subsequence (£„k) of coordinates which converges to some o with |cr| = l, then it can be divided into two such subsequences, and we can define w(1) and u{2) as weakly adherent points of the sequences e„2k_, and e„2k; as before we obtain two different supporting hyperplanes through x. If x has two coordinates of modulus 1, then there are clearly two supporting hyperplanes through x. Finally if at has the form aek + x\ with |<t| = 1 and ||£'||<1, then for sufficiently small t we have p{x + ty) = \(T-\-trjk\, so that the norm is weakly differentiate at x9 with p'(x,x)) = 5R(er*7k). The points of the form <jtk + x' with ||x'||<l are therefore the points of smoothness of the unit ball of /°°. 8. Suppose that © is a bounded domain in the complex plane, and that HB((&) is the (B)-space, defined in §14,9., consisting of functions
6. Uniform convexity 353 which are holomorphic in © and continuous on the boundary S. Since ||/(z)|| = sup|/(z)| = sup|/(z)|, HB((&) is norm-isomorphic to a closed ze© zeS linear subspace H of C(S). By 3.(4) and the first example of this number, all the functions f(z) which attain their maximum modulus 1 at just one point of S are points of smoothness of the unit ball. We shall assume that S is a homeomorphic image of the circumference of the unit circle. Then using results in the theory of conformal mappings, to each point t of S there is an f(z)eHB((&) which attains its maximum modulus at t. and at t alone. By §22,3.(4) the dual HB(&)' is norm-isomorphic to the quotient space yjl(S)/HL, where H1 is the space of all measures which vanish on H. By 3.(1)a) the points o8t, where |<r| = l and 8t is the coset in yjlfiyH1 corresponding to the point measure 8t, teS, are extreme points of the unit ball K' of SHl(S)IHL. We shall show that these are the only extreme points of K'. Since the unit ball K' of $R(S) is weakly compact, its image under the canonical mapping from $R(S) onto 9M(S)//f-1 is also weakly compact, and so it is equal to K'. It then follows from §25,1.(9) and §25,2. that the points <rdt are the only possible extreme points of K'. If we now apply 3.(1) a) again, it follows that the points of smoothness of the unit ball of HB(©) are precisely those f(z) which attain their maximum modulus 1 at just one point of S. The extreme points of the unit ball of HB{<&) are more difficult to determine. For the case where 05 is the unit disc in the complex plane cf. de Leeuw and Rudin 6. Uniform convexity. Following Clarkson [1], a normed space E, and its closed unit ball, are said to be uniformly convex if for each c with 0<c^2 there exists a <5(e)>0 for which it always follows from ||x||^l, ||y||^l and ||x-j;||^£ that ||£(x + j0||^l-<5(e). Such a function 8(e) is called a module of convexity for £. By 1.(1) f) every uniformly convex space is strictly convex. (1) E is uniformly convex if and only if it always follows from \\xn\\ g 1, ILyJ^l and lim Hi(^cM + ^JII = 1 that l\m\\xn-yn\\ = 0. n-* oo n~* oo Proof. The condition is necessary, for a contradiction of uniform convexity follows immediately from ||x„k — y„J^;£ and lim^(x^4-^)11 = 1. n-* oo On the other hand, if the condition were satisfied, and if for some 8>0 there were no 8(e) with the required properties, then there would be sequences x„, yn with ||y(x„ + 3;J||^l and ||x„ — yj^e, which would contradict lim||xn — yj = 0. 23 Kothe, Topological Vector Spaces I
354 § 26. Metric properties of normed spaces (2) Every linear subspace and every quotient space of a uniformly convex normed space is again uniformly convex. If E is uniformly convex, so is its completion. We need only prove this for quotient spaces E/H. Suppose e>0 is given, and suppose that ||x||^l, ||j)||^l and ||ic —j)||^e. Then there exist xex, yey with ||x||^l+A, \\y\\^ 1 + A, A>0, with A>0 arbitrarily small. It then follows from ||x —j/|| ^ ||x —j>|| ^ e that 2(* + )>) < 1 < \-s 1+A • (l + A). Consequently lim S a-o+ Vl+A c ^ <51 — I is a module of convexity for E/H. (3) Suppose that E is a uniformly convex normed space, that u0eE' and that ||w0|| = l. // x' and x" are two elements of E, with ||x'||:gl and 8(e) ||x"||^l, and if they satisfy the inequality \uQx —1|< , then ||x'-x"||<;;. This follows directly from the inequality IlKx' + x'Oll^iMx' + x'Ol^lMoxV^M^'-^^l-^) and the assumption of uniform convexity. If x0 is a point of the unit sphere ||x|| = l in E, and if w0 is chosen in such a way that u0x0= 1, then (3) means that the weak neighbourhood S(c) of x0 in the unit ball K given by \u0x — 1| < is contained in the strong neighbourhood of x0 in K given by ||x0 — x||<c. Thus the weak and the strong topologies coincide on the unit sphere of a uniformly convex space. It should be observed that it does not follow from this that the corresponding uniform structures are the same; the unit sphere, being bounded, is certainly weakly precompact, but in general it is not strongly precompact, as for example the uniformly convex space I2 shows (cf. the next number). We now obtain Milman's [1] important criterion for reflexivity: (4) Every uniformly convex (B)-space is reflexive. Proof, (cf. Dieudonne [2]). If is sufficient to show that every element zeE" with ||z|| = 1 is contained in the closed unit ball K of E. z belongs to the polar K°° of K° in E", and so, by the theorem of bipolars, it is a weak closure point of K. As V runs through the weak neighbourhoods of z in E", the sets Vn K form a weak Cauchy filter g on K with limit z. If we can show that sets of arbitrary small diameter, in the sense of the norm on E, are contained in this filter, then g will be a Cauchy filter
7. The uniform convexity of the lp and LP spaces 355 with respect to the strong topology on K; since K is strongly complete, g will therefore have its limit in K, so that z will belong to K. Since ||z|| = l, given d>0 there exists a u0 with ||m0|| = 1 and S S \u0z —1| < —. The set W3 of all xelC with \u0(z — x)| < — belongs to S the filter g, and if xeWd then |m0x —1| < —. If d = (5(e), then it follows from (3) that ||x' —x"|| <e for x', x"eWd; the diameter of Wd is therefore at most c. (4) can be generalized to arbitrary locally convex spaces: (5) A sequentially complete locally convex space E\%\ is semi- reflexive if every bounded subset of E is contained in a bounded closed absolutely convex uniformly convex set B. Proof. By §20,11.(2), EB is a uniformly convex B-space, and X defines a weaker topology than the norm topology of EB. By (4), EB is reflexive, so that B is ^(E^-compact. Since every ueE' defines a ^-continuous linear functional on EB, ZS{E') is coarser than £s(£y on EB; B is therefore ^Xs(£')-compact. The assertion now follows from §23,3.(1). 7. The uniform convexity of the lp and IF spaces. The Hilbert space I2 is easily seen to be uniformly convex. In I2 we have ||i||2=ii, and from this follows the identity (1) l|3t + ^l|2 + p-^ll2=2(||3t||2 + ||t)||2). II 1 P Thus if ||*||^1, ||i)||^1 and \\x-x)\\^e, then that I2 is uniformly convex. The proof of the next result is also comparatively simple. (2) lp and LP are uniformly convex for p ^ 2. This depends upon the following inequality, valid for all complex a, fi and for p^2: (3) |a + i?|p+|a_i9|p^2^1(|ar+|/ir). Proof. Using (1) it follows from § 14, 8.(9) that (|a + ]8r + |a-i8|r^(|a + i8|2 + |a-i8|2)Y = l/2(|a|2 + |i8: 2 p-2 Using Holder's inequality for —| = 1, we get P P 2 p—2 2 p—2 M2 + |/?|2^(|a|'+ |/W1 + 1)~= (|a|'+ |j8|')^2~. 1 < 1 , so 2 |\2 23*
356 § 26. Metric properties of normed spaces Combining these, we have 1 p-1 i (|a + /?r+|a-/?|')p^2 p (|a|'+|j8|')p, which gives (3). By summing over the components of vectors in lp or integrating over functions of LF, we obtain the relation (4) l|x + ^||?+||x-^||^2^-1(l|x||^+|b||^), p^2, from (3), for any two elements of lp or LF, respectively. Now if ||xjp^l, |bJ|p^l, lim||xB + jJp = 2, then it follows from (4) that lim||x„ —yjp = 0, and uniform convexity follows from 6.(1). The case l<p<2 is substantially harder to deal with, as we shall shortly see. It follows from (2), 6.(4) and § 14,10. that the Lp-spaces are reflexive for p^2. Using this, we can now derive a result annonced in §14,10.(14): (5) If \<p<oo and, - + -= 1, the spaces U and 13 are each the dual of the other. ^ ^ It is sufficient to show that (U)' = I3 for p^2. For then (I3)' = LF for l<q^2, since LF is reflexive. If g(t) is an element of 13, and if we define a linear functional for all b fell by setting (g,f} = $f(t)g(t)dt then it follows from Holder's b a inequality that j \f(t)g(t)\dt^ \\f\\p \\g\\q, so that <#,/> is continuous a on LF. So 13 can be identified with a subspace of (LF)'. We have (6) sup |<0,/>|=||0||,. Il/ll^i Holder's inequality implies ^ in (6). If we put f(t)=\g(t)\q~1 e,g(t), with 9(t) c«W = -r7-T for 0(f)#0 and sJt) = 0 for g(t) = 0, then fell, since \g(t)\ \\f\pdt = \\g\p{q-l)dt = \\g\qdt=\\g\\\. It is immediately verified that ' TmT/ = "^"9' which establishes (6)- This means that on 13 the norm of (LF)' coincides with the norm of 13. Since 13 is norm-complete, 13 is therefore a closed linear subspace of (Lp)'. If 13 were a proper subspace, there would be a continuous linear functional on (Lp)' which vanished on 13 without vanishing on the whole of (Lp)'. Because LF is reflexive, this functional must be given by an feLP. But it follows from §fgdt = 0, for all geI3, that in particular, putting g=\f\p~1sfeI3, we obtain \\f\pdt = 0. This means that/ is the zero element in LF, which gives a contradiction.
7. The uniform convexity of the lp and LP spaces 357 The usual proof of (5) uses the theory of differentiation of real functions, for example the Lebesgue-Nikodym theorem (cf. Banach [3] p. 61 ff or Bourbaki [7] Vol. 2 p. 55, for example), while the present proof makes do with the results already used in § 14,10. It follows from the remarks of § 14,10. and the proof of (5) that (5) is also valid for LP(— oo, + oo). The dual of L1 can now be determined, as well. (7) The dual of L1 is L°°. The underlying interval of the real line may again be either \a,b\ or (—oo, +oo). We first investigate the case L1[a9b']. Once again, L00 c= (L1)'. Suppose that u is a continuous linear functional on L1. By § 14,11.(4), L1 => L2, and u is also a continuous linear functional on L2. By (5), there is therefore a function g(t)eL2 with (u,f} = $gfdt for all feL2. If M is a measurable subset of \_a,b~\, and if cp(t) is its characteristic function, then cq(t)(p(t) belongs to L2, and we therefore have {u,cgcp} = J \g\dt^\\u\\fi(M). But it follows directly from this that esssup|#|rg||w||, M i.e. gel?. The formula (u,f} = $gfdt therefore holds for all feL\ since L2 is dense in L1. The case Lx(— oo, + oo) can be reduced to this one: L1 [ — n, n\ can be considered as a closed subspace of L*(— oo, + oo). On Lx[ — rc, n\ the continuous linear functional u is determined by a function gn(t) which is essentially bounded by ||w||. These functions gn(t) are the restrictions to [ — n, n\ of a function g(t) defined on ( — 00,00) and essentially bounded by ||w||, and this functions satisfies <m,/> + 00 = j g(t)f(t)dt. The exact details of the proof are left to the reader. - 00 We shall now show that lp and LP are uniformly convex for 1 <p<2, as well. We follow a method due to McShane [1], which works for all p with 1</?<oo, and which can be applied to more general classes of (B)-spaces. Theorem (12) was first proved by Clarkson [1]. (8) Suppose that E is a uniformly convex normed space, and that l<p<oo. Then given e>0, there exists a Sp(c)>0 such that it always follows from ||x||^ 1, ||y||^ 1 and \\x — y\\^c that (9) \(x + y) ^(l-<5p) MI'+I For arbitrary x, yeE, we therefore have (9') 1 (x + y) S M-5. Il*-3>ll PVsup(||x||, \x\\»+\\y\\^
358 § 26. Metric properties of normed spaces Proof. By 15, 9.(6) we have (10) (i(l+c))p^(l+c') for c^O. Since \+tp attains its minimum at t=\, we have strict inequality (i+ty in (10) for 0^c<l. It is-sufficient to prove (9) for ||x|| = l, \\y\\£ 1 and ||x-.y||^. We assume that (9) is false. Then there exists e>0 and elements x„, yn with \\xn\\ = \AynU\Axn-yn\\^ and (ID M(xH + yn)\\p H\M\p+\\yn\\p) l. We first show that this is only possible if ||yj->l. If not, there would be yn with HyJrg^K 1, with (11) holding. But it would follow from this and the remark following (10) that we would have 1 ■(xn + yn) ^ (|(i + lkll)Y^ |(i + lkllp) = f (IKir+lkD for all n and for some p < 1, which contradicts (11). yn If we now put zn = , then ||z„ —y„||->0. Consequently if n^n0, ■^(xn + z„) = 1. But by Ik ||x„ — zn\\ ^ —. Further, because of (11), lim 6.(1) this contradicts the uniform convexity of E. (12) lp and LF are uniformly convex for all 1 </?< oo. We first prove this for U. Suppose that /, g are in LP, and that ||/||pgl,y|p^l and ||/-0||p^e. /maybeeither [o,b] or (-oo, +oo). Let M be the set of all tel for which (13) |/(r)-0(t)l' ^ ^(\f(t)\p+\g(t)\p) ^ ^sup(|/(t)|Mflf(t)|") Applying (9') to E = K, we obtain the following inequality on the measurable set M: (14) 1 ■(f(t) + 0(t)) *li-*p[^))(j(\fW+MW)- On N = I~M we have j \f(t)-g(t)\"dt ^ ^ j (|/|"+ \g\")dt ^ |. N J Cp \f-g\pdt ^-. If fM and gM denote the It follows from this that
8. Further examples 359 restrictions of/ and g to M, we therefore have ||/m-#mIIp ^ ;^> so that sup(||/M||p, \\gM\\p) ^ (15) 2-21/p , i.e. supf ji/rdt, n^i^tU—T. It now follows from (14), (15) and the fact that the integrand of the first integral is non-negative, by (9), that J{ \{\f\p+\g\p)-(^\f+g\^}dt^ M M V, 4*/P \f\p + \g\p)dt >s, p\4llpJ 2P + 2 ' Finally it follows from this that 1 <f + 9) ^ IS, p\4l,pJ 2P+2 Up for all fg with ||/||p^l, \\g\\p^\ and ||/-flf||p^c. The proof for lp is similar; it is sufficient to replace the integral over / by the sum over n=l,2,... . Let us observe that these results are also valid for the spaces /J, where d is any cardinal, and that the proofs are the same. 8. Further examples.If En, n = 1,2,..., are (B)-spaces, let lp(En), with l^/?<oo, denote the space of all sequences x = (xn), xneEn, with 00 X llxJP< °°- Arguing as for the special case lp in § 14, 8., it is not diffi- «=i / * y/p cult to see that lp(En) is again a (B)-space, under the norm I £ \\xn\\p I The norm induced on En is identical with the original norm. If all the spaces En are equal to the same (B)-space E, we write lp(E). Similarly l*°(En) is the (B)-space of all sequences x = (xn\ xneEn, with ||x|| = sup||xJ<oo. 1 (1) The dual of lp{En\ p>l,is lq{E'n)l - + - = 1 1, and the dual of l\En) is r(E'n). \P « ' We shall prove this for p>\. Every u = (u„)elq{E'n) defines a continuous linear functional on lp(En\ as is clear from the inequality \ux\ = Z UnXn < ziMiixj^(Lii«jr,(iwip)i/p < 00.
360 § 26. Metric properties of normed spaces On the other hand every continuous linear functional must be of 00 the form ux= £ unxn, with uneE'n. For each x = (xn) we can find an n= 1 x = (xB) with ||xJ = |l*JI and unxn = \\un\\ ||xj-fin for arbitrary e„>0. It then follows from Xw„xn<oo that £||mJ ||xJ<oo. Since this holds for all xelp(En), we have \\\un\\)el\ so that uelq(En). Similarly, it is easy to confirm that the norm of lq{E'n) is the same as the dual norm of lp{En). It follows directly from (1) that (2) lp(En),p> 1, is reflexive if and only ij all the spaces En are reflexive. We now establish (3) lp(En) is strictly convex if and only if all the spaces En are strictly convex. If En is not strictly convex, neither is lp{En), for the boundary of the unit ball of En is a subset of the boundary of the unit ball of lp(En). Conversely suppose that all the spaces En are strictly convex, and suppose that x = (x„) and y = {yn) are two distinct elements on the boundary of the unit ball of lp{En). Then the vectors x = (||x„||) and x) = (\\yn\\) in IP also satisfy ||s||p = ||i,||p=l. If x±T), then ||±(s + i))||p<l, since lp is strictly convex. It follows from this that |li(x + j;)||<l. If x = x), there is at least one n0 with HxJIHIyJI and x„0*yno. Then lli(^o + ^o)ll<ll^oll» and aSain {t follows that |li(x + y)||<l. (4) lp(En\ p>l, is separable if and only if all the spaces En are separable. The simple proof is left to the reader. (5) lp(E) is uniformly convex if and only if E is uniformly convex. It is easy to see that the condition is necessary. On the other hand if E is uniformly convex the proof of uniform convexity can be followed through in exactly the same way as for lp, with the difference that 7.(8) must be applied to E, and not to K. Day [3] has shown that, more generally, lp(En\ p>l, is uniformly convex if and only if the spaces En have a common module of convexity 3(e). The generalization of the //-spaces to spaces of the form LP{E), which is essentially more important in applications, will be investigated in greater depth in the second volume. 9. Invariance under topological isomorphisms. The properties investigated in this paragraph are of a metric kind. The question arises of whether concepts can be obtained from them which are invariant under topological isomorphisms. Now two normed spaces are top-
9. Invariance under topological isomorphisms 361 ologically isomorphic if and only if they are norm-isomorphic under some suitable equivalent norms (cf. § 14, 2.). We therefore call a normed space E strictly, smoothly or uniformly normable if an equivalent norm can be defined under which E is respectively strictly convex, smooth or uniformly convex; in this way we define concepts with the required properties. Theorem 6.(4) can now be expressed in the stronger form: every uniformly normable (B)-space is reflexive. Day [2] has shown that the converse is not true; a counterexample is given by (1) lp{En), En = l™, p> 1, is reflexive, but not uniformly normable. Proof. /„*> is the space of all (£l5...,U with ||(£1,...,£II)|| = sup|6|. That E = lp(En) is reflexive follows from 8.(2). If E were uniformly normable there would be a second norm 11x11' on E with (2) ||x||^||x||'^M||x||, for which E is uniformly convex. It then follows from ||x||'^l, IMI'^l and \\x-y\\'^s that 5(e) =1-sup ||£(x + y)|r>0. Suppose that z = (a1,...,a„), |aj = — = a. Then z belongs to /„°°^E, and, since ||z|| = l/M, it follows from (2) that ||z||'^l. We set z = (a1,...,an_1,-aj. Then z-z = (0, ..., 0, 2a„) and \{z + z) = (a1,...,an_1,0). It follows from \\z-z\\'^\\z-z\\ = 2oc that ||£(z + z)|r ^ 1 — 8(2a). This means that a«-i K\-S(2a) l-8(2<x) If we now apply the same idea to ,0 <1. 1 and 1 l-8(2<x) 1 S/-I ;(ai,...,a„_2,an_1,0) 1 — o(2(X) (a1,...,a„_2, — an_l50), and so on, we finally obtain that \\yn\\' = v(l-(5(2a)r-1 1 Since on the other hand \\y„ 0,...,0 <1. — -—, and since (1 -(5(2a))""1 M(l-(5(2a))"-1 v v " tends to 0 as n->oo, we obtain the contradiction HyJ'^HyJ^l, for sufficiently large n. The uniformly normable (B)-spaces therefore form a proper subset of the collection of reflexive (B)-spaces.
362 § 26. Metric properties of normed spaces It is not known if every reflexive (B)-space is strictly normable. There are however non-reflexive (B)-spaces which are strictly normable, as we shall see. The situation is the same for smooth normability. (3) A normed space E is strictly normable if and only if there is a continuous one-one linear mapping A from E into a strictly convex normed space F. The norm ||x||'=||x|| + ||;4x|| is then a strictly convex norm on E, equivalent to \\x\\. Proof. The equivalence of the two norms follows from ||x||^||x||' ^(M + l)||x||, where ||i4|| = M. Suppose now that IWHIxJ^l and l|xi + X2ir = ll*iir + ll*2ir- This is only possible if ||x1+X2ll = ||x1|| + ||x2|| and ||y4x1+y4x2|| = ||/lx1|| + ||y4x2||. It tnen follows from the strict convexity of F and l.(l)g) that Axl=aAx2, with a^O, so that x1=<xx2. Since ||x1||/ = ||x2ir, this is only possible if xt=x2, so that E is strictly convex under ||x||\ It is trivial that the condition is necessary. (4) Suppose that E is smoothly normable. If E has a continuous one- one linear image in a space F which is simultaneously strictly and smoothly normable, then E is also simultaneously strictly and smoothly normable. We can suppose that E is smooth and that F is both smooth and strictly convex. By (3), the norm ||x||'=||x||+ ||v4x|| is strictly convex. But it is also smooth, for by 4.(13) both ||x|| and ||^4x|| are weakly differentiate norms on E, and so also is their sum ||x|| + ||/lx||. Applying 4.(13) again, it follows that ||x||' is smooth as well. (5) Every separable normed space E is simultaneously smoothly and strictly normable. The dual of every separable (B)-space is strictly normable. Proof. Let {xj be a sequence dense in the sphere ||x|| = l of E. A continuous one-one linear mapping from the strong dual E into (U X U X • \ —-^,...,—^,...1. If we again set ||w||t = ||m|| + ||,4w|| on E, then by (3) \\uWi is a strictly convex norm on E which is equivalent to ||w||. Thus the second part of the theorem is proved. The mapping A sends every weakly convergent sequence in E into a norm-convergent sequence in I2. It follows directly from this that the new unit ball Kx of E, given by ||w|| + ||,4w||^l, is weakly sequentially closed, and so it is weakly closed, by §21,10.(7). By taking K'^ as unit ball, therefore, we obtain a new norm Hxllj which is equivalent to the original norm and which gives {{u^ as the dual norm, since K1=K^\ \\x\\r is therefore smooth, by 3.(2).
10. Uniform smoothness and strong differentiability of the norm 363 If we now map E, with norm Hx^, into I2 by setting fu1 x utx \ Bx = \ ,...—j-,...), where ut is a weakly dense sequence in ||w||i = 1, then the theorem follows from (4). In this result of Day's [6], the separability condition cannot be relaxed. None of the spaces If with d>X0 is smoothly normable (cf. Day [6]). By (5), Z1 is both strictly and smoothly normable; its dual /°° is strictly normable, by (5), but is not smoothly normable, as Day [6] has also shown. 10. Uniform smoothness and strong differentiability of the norm. There is a further concept of convexity, which is dual to uniform convexity, and which provides a refinement of the concept of smoothness. A normed space, and its norm, are said to be uniformly smooth if for each £>0 there is an ^(e)>0 for which ||x||^l, ||y||^l and \\x-y\\^z always implies \\x + y\\ ^ ||x|| + ||;y||-e||x-j;||. (1) A normed space E is uniformly smooth if and only if for each e>0 there is a p(s)>0 for which ||x||=l, |[y||^p always implies that a) Suppose that E is uniformly smooth, and that there are sequences x„,y„eE with ||xj = l, |bJ->0 and ||xw + j;J + ||xb-jJ>2 + c0||j;J. Setting xn+yn = vn,xn-yn = wn, we obtain ||ij + ||wj>2 + —||i>n-wj. Since ||tf„ + wj = 2, this means that Ik + wJ^lkll + llwJ-^lk-wJ. The inequality remains valid when we replace vn and w„ by vn w„ v'» = -i—r~7 and < = -,—m> resPectlvely- Since IKII^i, KH^i and ||^ — w'n\\ ^ —, we obtain a contradiction of the uniform smoothness by letting n^co. b) Suppose that E satisfies the conditions of (1), and suppose that there exist c>0 and sequences xn,yn with ||xj|^l, ||j/J^l, ||xn —yJ->0 and ||x„ + )J<||xJ + ||)J-£||x„-)J. Setting xn + yn = sn, xn-yn = tn, we obtain \\sn + tn\\ + \\sn-tn\\>2\\sn\\ + 24tn\l For s'n = -\ and t'n = -\ this gives K + Q\ + K-Q\>2 + 2e\\Q\9
364 § 26. Metric properties of normed spaces and this provides a contradiction to the conditions of (1), since ||sj = l and IKII-0. As we saw in 4, smoothness is equivalent to the weak differentiability of the norm. Uniform smoothness can also be expressed in terms of differentiability. Once again let q(x) be the Minkowski functional of a closed convex £-body Cbo in a locally convex space E\X\. The function q(x) is said to be strongly differentiable at a point x0 if there is a continuous real linear functional u on £[£], and a monotonically increasing function SXo(p), p>0, with \imSxo(p) = 0, for which (2) \q(*o+y)-<i(xo)-uy\^q(y)$xo{q(y)) holds for all yeE. u is called the strong or Frechet derivative of q(x) at the point x0. If q(x) is a norm, (2) can be written more simply as (3) lim (\\Xo + y\\-\\Xo\\-uy) = 0. Ibll-o \\y\\ (4) // the strong derivative of q(x) exists at x0, then it is equal to the weak derivative q'(x0,y). The function q(x) is strongly differentiable at q(x0 + ty) — q(x0) x0 if and only if it is weakly differentiable at x0 and converges uniformly in q(y)^l to q'(x0,y). Proof. Uniform weak differentiability at x0 means the existence of a function Sxo(t) with lim Sxo(t) = 0 for which (5) \q(x0 + ty)-q(x0)-q'(x0,ty)\^\t\SX0{\t\) holds for all y with q(y)^ 1. But this is equivalent to (2). q(x) is said to be uniformly strongly differentiable in E[X] if q(x) is strongly differentiable at every point except the origin, and if there is a S(p) such that (2) (or (5)) is satisfied, with Sxo(p) = d{p), for every x0 with ||x0|| = l. (6) A normed space E is uniformly smooth if and only if its norm is uniformly strongly differentiable. a) Suppose that E is uniformly strongly differentiable. The inequality 4.(6) gives ,., \\x0-ty\\-\\x0\\ l|x0 + ^ll-IM (7) ^ q {x09y) ^ .
10. Uniform smoothness and strong differentiability of the norm 365 From (7) and (5) we obtain (8) l|xo + ^|| + ||x0-^||^2 + 2|t|5(|t|) for all ||x0|| = l and all |[y||^l; but by (1) this means that £ is uniformly smooth. b) Conversely if E is uniformly smooth, then (8) holds for some suitable function 6{p). By 4.(9) the weak derivative q'(x0,y) exists, and by arguing back from (7) and (8), we obtain uniform strong differentiability. In particular, we have also shown that every uniformly smooth normed space is smooth. (9) // the (B)-space E is uniformly convex, then E' is uniformly smooth. We assume that E is not uniformly smooth. Then by (1) there exist e0>0 and sequences u„, vneE' with ||wj = 1, ||t;J|->0 and (10) \\un + vn\\ + \\un-vn\\>2 + z0\\vn\\. Since E is reflexive, by 6.(4), there are elements xn, x'n in E with II* J = IK|| = 1, \\un + vn\\ = {un + vn)xn and \\un-vn\\ = {un-vn)x'n. Now |lk + t>J-l|^lkll, so that \(un + vn)xn-\\^\\vn\\; consequently \unxn-\\S2\\vn\\. Similarly \unx'n- 1| ^ 2||t>J. 8(e) Now if 2||uJ < it follows from 6.(3) that ||xw — xj<e. It follows from this that \K + vJ-\-\\un-vn\\=\(un + vn)xn + (un-vn)x'n\ S\unxn + unx'n\ + \\vn\\\\xn-x'n\\ ^2 + eK||, which contradicts (10), for c<c0. (11) // the (B)-space E is uniformly smooth, then E is uniformly convex. Suppose that E is not uniformly convex. Then by 6.(1) there are sequences un9 vn with ||wj<; 1, |k||^ 1, lim|k + uj = 2 and \\un-vn\\^c0. We may assume that \\un + vn\\ >2 . 4n For each n there exist x„ and x'n in E with ||xj = ||xj = l and {un + vn)xn^\\un + vn\\ --^-, (un-vn)x'n^\\un-vn\\ --^. 8rc 8n
366 § 26. Metric properties of normed spaces Then we have 1 x'«\ \xn + — 1 n 1 + \xn - _ x'n\ n \ ^un\xn + + vn \ xn = {un + vn)xn + (un-vn): 1 e0 ^ \\un + v„\\ +-\\un-vH\\ - — n 4n In n 2 + Since 0, this contradicts (1). It follows from (11) that every uniformly smooth space is reflexive, since its dual is. From this fact, (9) and (11) we obtain (12) A normed space is uniformly smooth (respectively uniformly convex J if and only if its strong dual is uniformly convex ( respectively uniformly smooth). The theorems of this number are due to Day [5], Smulian [3], [4], [6] and Bourbaki [6], Vol. 2 pp. 144-5. 11. Further ideas. Uniform convexity can be weakened in the following way: a normed space is said to be locally uniformly convex if to each e>0 and each xeE with ||x|| = l there corresponds a S(e,x) for which ||j(x + )>)||5n — <5(c,x) for all Hyll^l with Hy —x0||^c. This concept has been investigated in detail by Lovaglia [1]. If E is locally uniformly convex, then the norm of E is strongly differentiable. If E is locally uniformly convex, then E has a strongly differentiable norm, provided that the additional assumption is made that for each ueE there is an x in £ with ||x|| = l and mx = ||m||. Every reflexive space has this property. The question of whether conversely every (B)-space with the property that each ueE attains its supremum on the unit ball of £ is reflexive originated with Mazur. James [4] first answered this question in the affirmative for separable (B)-spaces, and recently (James [5]) gave the same answer for arbitrary (B)-spaces. Fan and Glicksberg [1], [2] have introduced and investigated further concepts of convexity which are weaker than uniform convexity. In this way they have obtained a stronger form of Eberlein's theorem for normed spaces. Cudia [2] has recently obtained important results by refining the convexity and differentiability properties further; these results complement and round off the results described here. The survey article of Cudia [1] gives a full account of the present state of the theory.
CHAPTER SIX Some Special Classes of Locally Convex Spaces If the topology of a locally convex space is the same as the strong topology, the space is said to be barrelled. If the topology is the same as the Mackey topology, and if every linear functional which is bounded on the bounded subsets of the space is continuous, then the space is said to be bornological. Every (F)-space has these properties, and each of these properties entails a sequence of important consequences. The investigation of the properties of barrelled and of bornological spaces, which generalises the theory of (F)-spaces, and which goes back to Mackey and to Bourbaki, forms an important part of the general theory of locally convex spaces. These two classes of spaces are considered in depth in § 27 and § 28. The theory of (F)- and (DF)-spaces which is given in §29 provides an application of these results. (DF)-spaces were introduced by Grothendieck to provide a class of spaces which contains the duals of (F)-spaces as a special case. In § 30, the theory of perfect spaces, which was originated by Toeplitz and the author, is developed as a special case of the general theory of locally convex spaces. The special structure of sequence spaces to some extent allows simplifications to be made, and to some extent enables further results to be established. Some of the questions of the general theory which have been left open until now are answered in §31 by giving counterexamples, which are perfect spaces. Besides this, § 31 contains a discussion of the problem of the existence of complements; although a collection of counterexamples is known, there are only a few general results on this topic. § 27. Barrelled spaces and Montel spaces 1. Quasi-barrelled spaces and barrelled spaces. As in § 21,2., a locally convex space E[%] is said to be barrelled if every barrel in E is a ^-neighbourhood of o. This is equivalent to Z = Xb(E). As a result, every weakly bounded subset of E' is relatively weakly compact, by §21,4.(4). Every (F)-space is barrelled (§21, 5.(3)). By §23,6.(4), a locally convex space E[Xk(E)] is barrelled if and only if its dual E is ^(^-quasi- complete. By §23, 3.(4), the strong dual of every semi-reflexive space is barrelled. As in § 23,4., E\%\ is said to be quasi-barrelled if every barrel in E which absorbs all the bounded sets of £ is a ^-neighbourhood of o.
368 § 27. Barrelled spaces and Montel spaces This is the same as saying that :X = 2ft*(£'), by §23,4.(3). The 2-equi- continuous subsets of E' are then the strongly bounded subsets of E'. Every metrizable locally convex space is quasi-barrelled (§21, 5.(3)). We now develop some further properties of barrelled and quasi- barrelled spaces. Every barrelled space is quasi-barrelled. In the opposite direction, we have (1) // a quasi-barrelled space is sequentially complete, then it is barrelled. For it then follows from the Banach-Mackey theorem that %b{E') and %b*{E') coincide on E. It follows from (1) and §21, 4.(5) that (2) The quasi-completion and the completion of a quasi-barrelled space are both barrelled. aic following results give further hereditary properties. Every locally convex hull of quasi-barrelled (respectively barrelled) spaces is again quasi-barrelled (barrelled). Suppose that E\_(£]=YJAa(Fa\_(£a]), where the Fa[:Xa] are bar- a relied. If T is a barrel in £[£], then, because of the continuity of the maps Aa, each set Ta = A(a~1)(T) is again absolutely convex, absorbent and closed, and is therefore a barrel in Fa. By hypothesis, Ta is a ^-neighbourhood of o in Fa. But then (cf. § 19,1.) since Tis absolutely convex and contains all the sets ^4a(TJ, T is a ^-neighbourhood of o in £, so that £[£] is barrelled. If Ba is a bounded subset of Fa[£a], Aa(Ba) is bounded in E[3T|. If T absorbs every bounded subset of £, Ta therefore absorbs every bounded subset of Fa. Arguing as above, it follows that if the Fa[2a] are quasi-barrelled, so is £[£]. Since a quotient space (E/H)[%~] can be considered as the locally convex hull K (£[£]), where K is the canonical map of E onto E/H (cf. § 19,1.), the next result is a special case of (3): (4) Every quotient of a quasi-barrelled ( respectively barrelled ) space is quasi-barrelled ( barrelled). In particular (3) holds for inductive limits and locally convex direct sums. Thus all (LB)-spaces and all (LF)-spaces are barrelled. (5) The topological product of quasi-barrelled ( barrelled ) spaces is again quasi-barrelled (barrelled). This is a direct consequence of § 22, 5.(3). There exist spaces which are complete, but are not quasi-barrelled. For, applying § 23, 6.(6) to the case where £ is a non-reflexive (F)-space, E'P^jE)] is certainly semi-reflexive, but is not reflexive, and so it is not quasi-barrelled (§ 23, 5.(1)). But E'[Xk(E)] is complete, by § 21, 6.(4).
2. (M)-spaces and (FM)-spaces 369 Every complete space which is not quasi-barrelled is a closed linear subspace of a topological product of (B)-spaces, by § 18. 3.(7), and this space is barrelled, by (5). A closed subspace of a barrelled space therefore need not be quasi-barrelled. Every (not necessarily closed) linear subspace of finite co-dimension is, however, barrelled (for the proof, cf. Dieudonne [9]). A topological projective limit of barrelled spaces need not be quasi- barrelled, for by § 19,9.(1) each of the above spaces E'\_Xk(E)] is the topological projective limit of (B)-spaces. A normed space is quasi-barrelled, but need not be barrelled, as the example of §21,5. shows. On the other hand there are normed spaces which are barrelled, but not complete: a sequence kn of positive integers n is said to have density zero if lim — = 0. Let E be the dense linear kn subspace of Z1 which comprises all x = (£i)ell whose coordinates l{ vanish except on a set of density zero. It is easy to verify that the Xs(£)-bounded subsets of E' ' = /°° coincide with the norm-bounded subsets of /°°. It follows from this that the norm topology of Ec/1 coincides with the strong topology X&(£'). The strong dual of a barrelled space need not be barrelled, for there exist (F)-spaces which are not distinguished (cf. §23,7.(1) and §31,7.). The strong dual and strong bidual of a barrelled space need not be complete, since Komura [2] has shown that there are (M)-spaces which are not complete. 2. (M)-spaces and (FM)-spaces Besides (F)-spaces there is a further class of barrelled spaces which is of particular interest. A barrelled space Ep] is called aMontelspace, or (M)-space, if every bounded subset of E is relatively compact. Every (M)-space is clearly quasi-complete. A normed space which is also an (M)-space is locally compact, and so by § 15, 7.(1) it is finite-dimensional; an infinite dimensional (B)-space is therefore never an (M)-space. On the other hand there are (F)-spaces which are also (M)-spaces (the space co is an example of this). We call these spaces (FM)-spaces. It follows from the definition and § 23, 5.(1) that (1) Every (M)-space is reflexive. The topologies %b(E) and %C(E) on the dual of an (M)-space coincide. (2) The strong dual of an (M)-space E[}£] is again an (M)-space. Proof. E'[Xb(E)] is barrelled, since it is reflexive. The bounded subsets of E' are the same as the ^X-equicontinuous subsets. By § 21, 6.(3) 24 Kothe, Topological Vector Spaces I
370 § 27. Barrelled spaces and Montel spaces these are relatively Xc(£)-compact, and so, since %b(E) and £c(£) are the same, they are relatively Xft(£)-compact. Thus the weak and the strong topologies coincide on the bounded subsets of an (M)-space. In particular, we have (3) Every weakly convergent sequence in an (M)-space is also strongly convergent, to the same limit. We have the following hereditary properties: (4) The topological product and the locally convex direct sum of (M)- spaces are again both (M)-spaces. The strict inductive limit of a sequence of complete (M)-spaces is again an (M)-space. Proof. It follows from 1.(3) and 1.(5) that the spaces are barrelled; and from the structure of bounded sets in topological products (§15, 6.(13)), locally convex direct sums (§ 18, 5.(4)) and strict inductive limits (§ 19,4.(1) and (4)), and from Tychonoff's theorem, it follows that every bounded set is relatively compact. As we shall see in § 31, 5., there are quotient spaces of (FM)-spaces and closed linear subspaces of their dual (M)-spaces which are not (M)-spaces. For arbitrary (M)-spaces, we can obtain a counterexample as follows: It follows from (4) that every space of countable degree is an (M)-space. Thus (pat®coy and its dual wcp^cpw are (M)-spaces. Once again let Hx and H2 be the closed subspaces of (pco@co(p and axpOxpa), respectively, which were introduced in §13,6. By §23,5., {cp<x> 0 w cp)/Hx is not semi-reflexive, and so it is not an (M)- space; further H2 is not reflexive, and so it is not an (M)-space, either. We further observe that H2 is semi-reflexive, by §23,3.(5), and so it is not barrelled, by § 23, 5.(1). H2 therefore provides an example of a closed subspace of a locally convex space E on which the strong topology on E does not induce the strong topology of H2. There are (M)-spaces which are not separable, for example cpd9 d>K0. However Dieudonne [13] has proved that (5) Every (FM)-space E is separable. Proof. By § 18, 3.(7) we can consider E as subspace of a topological oo product TT En of normed spaces £„, where the projection Pn(E) of E n= 1 in En is equal to £„, for each n. Suppose that E is not separable. If all the spaces En were separable, oo TT En would also be separable, and so would £, by §4, 5.(1). We may n= 1 therefore assume that Ex is not separable. There is then a bounded uncountable subset N of Ex, whose elements are all at a distance ^ 3 > 0 from each other. For each x^eN we can determine an xeE whose projection in E1 is equal to x1. Let M be the collection of these x. We put
2. (M)-spaces and (FM)-spaces 371 M = M1. There is then a proper uncountable subset M2 of Mx whose projection P2{M2) is bounded in E2 (E2 is the union of countably many bounded sets). If we continue in this way, we obtain a properly decreasing sequence M„ of uncountable subsets, with each Pn(Mn) bounded in En. Suppose that xkeMk+l~Mk. The sequence Pnxk, fc=l,2,..., is then bounded in £„, since xkePn{Mn) for k^n— 1, so that xk is bounded in E. Since E is an (FM)-space, x„ has a subsequence which is a Cauchy sequence in E. The same holds for the projections ^(xj in Ex. But this is a contradiction, since the Px (xn) are pairwise at a distance ^ (5 apart. Likewise it follows from (2), (3), (5) and § 21, 3.(5) that (6) The dual of an (FM)-space is strongly sequentially separable. We have the following criterion (cf. Dieudonne and Gomes [1]): (7) A separable (F)-space E[}t] is an (M)-space if and only if weak and strong sequential convergence coincide in E'. Proof. Necessity follows from (2) and (3). Conversely suppose that E is separable, and that every weakly convergent sequence in E' is strongly convergent. Since E is barrelled, every weakly closed bounded subset M of E' is weakly compact. Since E is separable, the weak topology on M is metrizable (cf. §21, 3.(4)); by §4, 5.(4), M is therefore weakly sequentially compact, and so it is strongly sequentially compact. By §5,6.(3) every strongly sequentially compact set is strongly precompact, so that M is strongly precompact. Finally if we apply § 21, 6.(3) to the bounded closed subsets B of EaE\ it follows that each such B is relatively strongly compact in E". But B is strongly complete, as it is a closed subset of the (F)-space E\%\ and so B is strongly compact. E is therefore an (M)-space. The strong duals of (FM)-spaces can be characterised more simply. A locally convex space E\%\ has a countable fundamental system of absolutely convex compact sets if there is a sequence K1 a K2a ••• of absolutely convex compact sets such that each absolutely convex compact subset of £ is contained in some Kn. Following Dieudonne [15], we have (8) A barrelled space E\%\ is the strong dual of an (FM)-space if and only if E[X] has a countable fundamental system of absolutely convex compact subsets. Proof. If {£/„} is a base of absolutely convex neighbourhoods ofo in the (FM)-space £[£], the sets U° form a fundamental system of bounded subsets of E'[Zb(E)~\. By (2), the sets U° are absolutely convex and compact, and further E \%^\ is barrelled. The conditions are therefore necessary. 24*
372 § 27. Barrelled spaces and Montel spaces Conversely suppose that E[X] is barrelled, and that Kx a K2 <= ••• is a countable fundamental system of absolutely convex compact subsets of E[X]. If :X0 is the topology of uniform convergence on the absolutely convex compact subsets of £, £'[jX0] is metrizable, and so it is quasi- barrelled. Since <£ = Xb(E'\ the bounded subsets of £ coincide with the :X0-equicontinuous subsets, and so coincide with the subsets of the absolutely convex compact subsets of £[£]. £(jl]is therefore an (M)- space, and %b(E) and :X0 coincide on E'. E'\%q\ is therefore an (M)- space, by (2). Since £'[£0] is then quasi-complete and metrizable, E'[X0] is an (FM)-space. The assertion now follows from the reflexivity of E[Z] (cf.(l)). If E[X] is barrelled and if £[£] has a countable fundamental system of compact subsets, then £[£] is a dense subspace of the strong dual of an (FM)-space (Dieudonne [15]). Recently Garling [1] proved that E[X] is in fact the strong dual of an (FM)-space. This result and (8) have analogues for reflexive spaces (Garling [1]). 3. The space //(©). Some of the most important spaces of analysis are (M)-spaces, e.g. the spaces of infinitely differentiable functions, and of distributions, on a compact interval in P" (cf. Schwartz [1]). A detailed investigation of these spaces will be made in the second volume. Here we discuss a class of examples which motivated the name "Montel space". Suppose that (5 is a domain, i.e. an open connected proper subset of the complex sphere Q. The holomorphic functions x(z) defined on © which, in the case where © contains oo, vanish at oo form a complex vector space //(©). For every compact subset ft of © we define the norm (1) pa(x) = sup|x(z)|. This system of norms defines a locally convex topology £ on //(©), the topology of uniform convergence on the compact subsets of ©. If ©x c ©2 c ••• is a sequence of open connected subsets of ©, with 00 ©„<=©„+i and (J ©„ = ©, then X is also defined by the countable n= 1 collection of norms (2) Nl„ = ?*„(*) so that //(©) is metrizable. © need not be simply connected, but the ©„ can be chosen in such a way that each is bounded by finitely many simple closed rectifiable curves. This system of curves is denoted by Cn.
3. The space //(©) 373 By Weierstrass' theorem, a X-Cauchy sequence in H((5) always has a holomorphic function as limit (vanishing at oo), so that H((5) is an (F)-space. In addition, we have (3) //(©) is an (FM)-space. Proof. A subset B of //(©) is bounded if and only if px(x)^M($i) for all xeB. If (5 is a domain which does not contain the point oo, (3) is just the statement of Montel's theorem. The case where ooe(5 can be reduced to this by means of a transformation y(z) = x\ I, a$®. \z-aj By 2.(2), the strong dual of //(©) is an (M)-space. This space has a concrete representation, which we shall now derive. Suppose that (u,x} is a continuous linear functional on //(©). Then for a suitable choice of n we have (4) |<u,x>|^M||x||„. Now ||x||„ is the norm of the (B)-space HB((&n) of all functions which are analytic on (5„ and continuous on (5„ (cf. § 15, 9.). By the Hahn-Banach theorem, u can be extended from //((£>) to H £(©„), with the same bound. Consequently u is defined for all functions , AeO„ = Q~©„, A — Z which, considered as functions of z, are elements of HB((5n). By defining (5) U^=\U>JTZ we obtain a function of a defined in D„ which, following Fantappie, is called the indicatrix of u. If a=co, \/(a — z) is put equal to 0, so that w(oo) = 0, if oogO„. We observe that D„ is open, but that it need not be connected. (6) The indicatrix u(a) is locally holomorphic in D„ = Q~©„. If A0=Noo, then for a, A0e£)n we have U(l)-U(A0) I 1 ( 1 1 A — /iq \ A — /Lq \A — Z 1 -1 Now the difference quotient of converges to with A — Z (X0 — Z) respect to the norm of HB((&n). Since u is continuous, we therefore have u'(a0) = ( u, -z i. Thus the indicatrix is analytic at /0. \ (A0-z) V
374 § 27. Barrelled spaces and Montel spaces If /l0=oo, u(a) is analytic in a neighbourhood of oo, and u(a) con- 1 verges to 0 as /->oo, since converges to 0 with respect to the norm X — z of H£(©„) as /l->oo. #(/) is therefore also analytic at oo, and vanishes there. Conversely, the linear functional <w,x> on //(©) can be expressed in terms, of the indicatrix: (7) // u(X) is the indicatrix ofu in £)„, then (8) (u,x} = q) u(t)x(t)dt 2ni J C„+i /or each xe//(©). Proof. For each (possibly infinite) ze®„ and each x(z)eH((5) we have x(z) = — q) dr. 2 7Cf J t — Z The system of curves Cn+l is here described in such a way that (5„ 1 _x(^>)AtJfc) always lies to the left. We consider a sequence xk(z) = > —^ 27cf j ty — z of approximating sums for the integral x(z). Since there is a strictly positive least distance d between the points z of (5„ and the points of Cn+1, the sequence xfc(z) converges to x(z) uniformly for all ze(&n. Since u is continuous on HB((&n\ it then follows that <tt,x>=lim^X^f)\^lfcr—) ^^) = ^ I ^W^)^. fc-oo27ci \ ty — zf 2ni J C„+i We observe that £)„ is an open set containing 2I = Q~(5 and that On comprises finitely many components, each of which contains at least one point of 21; a set with these properties will be called an open neighbourhood H of 21, for short. Now suppose that #(/) is a locally holomorphic function in U, which vanishes at oo if oo lies in U. There is then an n for which Q~Uc ©„+i holds. The right hand side of (8) is defined for all xe#(©), and, as is easy to see, it defines a continuous linear functional <w,x> on H(G). The value of (8) is independent of n, provided that Cn+l lies in U; this follows directly from Cauchy's theorem. For the same reason, two locally holomorphic functions, defined on neighbourhoods Ul and H2 respectively, define the same linear functional if they agree on a neighbourhood of 21 contained in l^ n U2.
4. (M)-spaces of locally holomorphic functions 375 We say that two such functions are equivalent, and call a class of equivalent functions (vanishing at oo) a locally holomorphic function on 91; we also use u(a) to denote the class to which u{a) belongs. The set of all locally holomorphic functions on 91 forms a vector space We have already seen that every ueH(($))' is given by a usH(9l), and that conversely every ueH(tyi) defines a ueH((S)\ using (8). This last correspondence is clearly linear. In order to show that H(($)' and H(9l) are isomorphic, we must show that only one equivalence class u corresponds to each u. Now u{a\ considered as a locally holomorphic function on U zd 91, is determined by the values of its derivatives at one point in each of the components of U. But by hypothesis each component of U contains at least one point a0 of 9l = Q~(5. Thus it is sufficient to show that if the continuous linear functional defined by (8) vanishes identically on //(©), then u(a), together with all its derivatives, vanishes on 91. If ag91 and A 4= oo, then all the functions x(z) = , n=l,2,..., lie in //(($). Substituting these in (8), we get (z-Xf u{"\a) = 0 for n = 0,l,.... If /1=oog91, then the functions x(z) = z", n = 0,l,..., lie in H((5), and (8) shows further that all the coefficients of the Laurent expansion of u{a) at oo vanish. We have therefore proved the following theorem: (9) The dual of H(<&) is isomorphic to the space H(9I) of locally holomorphic functions u(a) on 91 = Q ~ 05, and the continuous linear functional u corresponding to a m(a)gH(9I) is given by (8). The last part of our proof shows that, in spite of the fact that we used the Hahn-Banach theorem to construct the indicatrix u{l), this is uniquely determined, up to equivalence, by the linear functional ue//(©)'. 4. (M )-spaces of locally holomorphic functions. It is natural to consider an arbitrary proper open subset O of Q, instead of a domain 05. O consists of at most countably many open components §f, and we again denote the space of all locally holomorphic functions x(z) on O by H(D). Clearly H(D) is equal to the Cartesian product of the spaces //(§f), and we define the topology £ on H{D) to be that of the topological product of the H(§f). It follows directly from this, using 2.(4), that H{D) is also an (FM)-space. It is easily confirmed that X coincides with the topology of uniform convergence on the compact subsets ft of O. In the present case, of course, the expressions 3.(1) defining X need only be semi-norms. The sequence (Sj cz (52 c "• can be chosen exactly as in 3., the only difference
376 § 27. Barrelled spaces and Montel spaces being that the set (£>„ can no longer be assumed to be connected; instead it can be taken to consist of just one connected component in each of §!,...,§„. This has the consequence that the Banach spaces HB(($>„) of locally holomorphic functions are of the type already introduced in §14,9. If 3.(4) holds, then u vanishes on all functions x(z) which vanish on §!,...,§„. Then //(§! u--u§J cz //£(©„), and we obtain the indicatrix of u as in 3. The evaluation of the dual can now be carried through as in 3.: a functional ueH{£))' satisfies an inequality 3.(4), and so it vanishes identically on Hi (J §f 1. If we consider u as a linear functional on V n + 1 / H\ [j §>il then Hi (J Jr>f I c HB((&n\ and we obtain the representation 3.(8) on Hi (J §t). This is however valid on the whole of H(O) = h( (J &)©//( (J &], since (8) vanishes identically on h( (J 9>\ The fact that the correspondence between the functional u and the locally holomorphic functions u(a) on 9l = Q~€) is one-one is obtained as in 3. We therefore have the following general result: (1) Using 3.(5) and 3.(8), the dual of H(£)) can he represented as the space H(tyL) of locally holomorphic functions on ^l = Q^O. If we give H(W) its strong topology lfc(/f(0)), H(ty) becomes a complete reflexive (M)-space by 2. and § 21,6.(4). Our aim is now to give a simpler definition of the topology ^(//(O)), which will enable the structure of the space H(ty) to be seen more clearly. Let ^l be an arbitrary proper closed subset of Q. Its complement 0 = Q~^l is then an arbitrary proper open subset of Q. Let Ox zdC2^>"- 00 be a sequence of open neighbourhoods of 51 with f] On = $l. We recall n= 1 that the sets On each have only finitely many components, each of which contains at least one point of 91. We further suppose that the boundary of On is a system Cn of finitely many rectifiable closed curves. 00 If ($„ is the sequence determined above for 0= (J §,., we can take For each €)„ we form the (B)-space HB(DJ of functions which are locally holomorphic in €)„ and continuous on On. We now identify each function in H B(D„) with its class of equivalent functions with respect
4. (M)-spaces of locally holomorphic functions 377 to 21, in other words, with a locally holomorphic function on 21. In this way HB(Dn) becomes a linear subspace of HB(£)n+1)9 and if(21) becomes the union of the spaces H B{£)n). It is now natural to try to interpret H(2l) as topological inductive limit of the spaces HB(On). Since the norm ||w|ln+i— sup |w(/)| of HB(£)n+ ±) induces a AeO„ + i _ coarser Hausdorff topology on HB(C„) than the norm ||m||„, we can 00 define the hull topology 2 on H(2I)= (J HB(On). We still have to show that "=1 (2) The hull topology on H(2l) is Hausdorff. First we obtain a simple lemma. (3) Every bounded subset M of HB(£)n) is relatively compact in H5(6„+1), and its closure M in HB(£)n+1) is compact in each of the spaces HB(On + m)9 mg; 1. M consists of functions which are uniformly bounded on On+1; by Montel's theorem, M is therefore relatively compact in HB(£)n + l). The rest of (3) follows from § 3, 2.(6). In order to prove (2) we must show that for each non-zero w0gH(21) there exists a ^-neighbourhood of o which does not_contain u0. Without loss of generality we may suppose that ii0eHB(O1). There is then a closed ball K[ about o in HB(C2) which does not contain u0. If Kx is the ball of equal radius in HB(Bi)9 u0 does pot belong to the closure Kx of Kx in HB(62), since K1 cz K\. By (3), Kx is compact in HB(5„)9 n^2. Since u0 does not belong to Kl9 there is a closed ball K'2 about o in HB(£)3) such that u0 does not belong to \~(K1uK'2). Again, let K2 be the closure in H B(£)3) of the ball K2 of equal radius in H B(£)2). u0 does not belong to \~(K1uK2), and this is compact in HB(£)n)9 n^3, by (3) and §20,6.(5). Repeating this procedure, we obtain a Z- n _ neighbourhood \~ Kt which does not contain u09 since u0 belongs to oo _ i = 1 no r K(. i= i The topological inductive limit lim H B(£)n) therefore exists, and it defines the hull topology *% on H(2l). (4) H(O) is the dual of H(2I)[2]. Proof. If <x,m> is a continuous linear functional on //(2(), we 00 have |<x,m>|^M for a suitable ^-neighbourhood (7= P Kn9 where n= 1 each Kn is a closed ball in H B(£)n). The linear functional x is bounded on the uneKn9 so that x is continuous on each HB(£)n).
378 § 27. Barrelled spaces and Montel spaces If, corresponding to 3.(5) (but with the roles of X and z interchanged), we introduce an indicatrix for x by writing (5) x(z) = ^x,^-y, zeS, then x(z) is locally holomorphic in each (&n = Q~£)n and so it is locally holomorphic in (J (£>„. If ueHB(£)n), we have n= 1 (6) <x,m> = (D x{t)u(t)dt. 2ni J The system of curves C„+1 is described in the same sense as in 3.(8), so that the domain ®n+1 is on the left. Conversely, using (6), every x(z)eH(£)) defines a linear functional on H(W\ provided that n is suitably chosen for each u. This linear functional is ^-continuous; for if u belongs to the neighbourhood Kn In of all u with ||m||w ^ — —— , where ||x||w+1 = sup \x(z)\ and IQ+ilNL+i zeC"^ |Cn+1| is the total length of the system of curves C„+l9 then |<x,w>|^l, 00 by (6). Thus |<x,m>|^1 through the whole ^-neighbourhood |~~ Kn. n= 1 (7) The hull topology % on H{^X) is identical with the strong topology Zb(H(Z»). By (4), % is a locally convex topology on H(tyL) for which the dual is H(O). By the Mackey-Arens theorem (§21,4.(2)), Xb(H(X))) is the finest topology with this property. (7) will therefore be established when we have shown that £ is finer than £fc(/f (O)). Let B be a bounded subset of //(O). There exist numbers M„^0 such that |x(z)|^M„ for all xeB and ze(&n. If now Kn is the set of all In ueHB(£)„) with ||m||w ^ , then |<x,m>|^1 for all ueKn Mn+\ \Cn+ j| 00 and all xeB, by (6); thus the ^-neighbourhood r~ Kn is contained in the strong neighbourhood B° of o. In particular if O is the whole complex plane l~, we obtain the space H(V) of entire functions. Its dual is H(co\ the space of functions holomorphic at oo. In the case where O is an open circular disc, this duality, and the results relating to it, were first obtained by Toeplitz [2] in the setting of the theory of perfect spaces (cf. § 30). For the general case, and for further generalisations, see Grothen- dieck [8], [9], Kothe [11], [12], Sebastiao e Silva [1], [3], Silva Dias [1] and TlLLMANN [1] tO [4],
1. Definition 379 § 28. Bornological spaces 1. Definition. A linear functional on a normed space is continuous if it is bounded on the unit ball. This can also be expressed by saying that every linear functional on a normed space which is bounded on the bounded sets is continuous. Expressed in this form, this property need no longer hold for arbitrary locally convex spaces, as we shall presently see. If we say that a linear functional we£* is locally bounded when its values remain bounded on any bounded subset of £, then the problem is to characterise those locally convex spaces for which every locally bounded linear functional is continuous. For this we can always suppose that the topology X is the Mackey topology, since we are concerned with a property which depends only upon the dual pair <£',£>, and not upon the original topology X. Following Bourbaki, a locally convex space £[£] is said to be bornological if every absolutely convex set M which absorbs all the bounded sets of £[£] is a ^-neighbourhood of o. Comparing this definition with the characterisation of quasi-barelled spaces given in § 23, 4.(3), we see that a definition of quasi-barrelled space is obtained from the definition of bornological space by adding the further requirement that M be closed. Thus we have (1) Every bornological space is quasi-barrelled. It therefore follows from § 27,1.(1) that (2) Every sequentially complete bornological space is barrelled. The topology X of a bornological space always coincides with Xk{E'\ and indeed with Xh*(E*\ since this is so for quasi-barrelled spaces (§ 27,1.). Our original question is answered by (3) A locally convex space E[X] has the property that every locally hounded linear functional on E is continuous if and only if E[Xk(E'\\ is bornological. a) Sufficiency. If E[X~] is bornological, X = Xk(E'), as we have just seen. If wg£* is locally bounded, and thus is bounded on every bounded subset B of E[X~\9 then the absolutely convex set M consisting of all xeE with |mx|^1 contains a scalar multiple of each set B. It is therefore a ^-neighbourhood of o, so that u is continuous. b) Necessity. We consider the locally convex topology Xx on E which is obtained by taking the absolutely convex sets which absorb every bounded set as a base of neighbourhoods of o. Xx is clearly finer than Xk(E'). By hypothesis, every Xx -continuous linear functional is
380 § 28. Bornological spaces continuous, and is therefore an element of E'. It now follows from the Mackey-Arens theorem that %* is coarser than %k{E'). The fact that <%k(E') = <Xx implies that £[£J is bornological. This characterisation of bornological spaces was used as a definition by Mackey [5] (who called them "relatively strong with a boundedly closed linear system"). Not every locally convex space £[£] with the property that (X = Zk is bornological, for not every such space is quasi-barrelled (cf. the example in §27,1.). The class of bornological spaces is however very extensive, as we shall see. As a first result, we have (4) Every metrizable locally convex space £[£] is bornological. Before proving this, we establish the following lemma: (5) // x„->o in a metrizable locally convex space £[£], there exist positive numbers p„, with p„->oo, such that pnxn->o as well. Proof. Let U1 zd U2=>"' be a base of neighbourhoods of o in £[£]. For each k there exists an nk>nk_1 such that xne—Uk for k n^nk. We obtain the required sequence by setting pn = k for nk^n <Wfc+l- Proof of (4). Suppose that u is a locally bounded linear functional on £, and that x„->o. Then pnxn->o as well, so that \u(pnxn)\ = pn\uxn\^M. But it follows from this that u x„->0, so that u is continuous. The assertion now follows from (3). It follows from (4) that not every bornological space is barrelled, since the example of § 21. 5. gives a normed, and therefore bornological, space which is not barrelled. Conversely, not every barrelled space is bornological, as Nachbin [4 ] and Shirota [1] have shown. 2. The structure of bornological spaces. We are also led to bornological spaces by asking another question. Suppose that £[£] is locally convex. Let us consider the finest locally convex topology £x which has the same bounded sets in E as does X. Its absolutely convex neighbourhoods of o must absorb all the bounded sets of £[£]. On the other hand, the collection of all the absolutely convex sets which absorb all the bounded sets forms a base of neighbourhoods of o for a locally convex topology on £, and this must be the required topology £x. £pXx] is clearly bornological, and the dual of £[£x] consists of all locally bounded linear functional on £[2]. The locally convex space £[£x] which is uniquely defined in this way is called the bornological space associated with £[£]. Clearly we have
2. The structure of bornological spaces 381 (1) E[X] is bornological if and only if E[X] coincides with its associated bornological space £[£*]. For each bounded, closed, absolutely convex subset B of a bornological space E[X] let us form the normed space £Bc£[I], as in § 20,11. % induces a coarser topology on EB than the norm topology. If we form the locally convex hull E[(£'~] = Y,Eb °f a'l these spaces EB, B then the hull topology X is finer than % by the definition of X. But on the other hand every T-neighbourhood of o U = \~ pBB, Pb>0, is also B a ^-neighbourhood of o, since it is absolutely convex, and absorbs all the bounded subsets of E[X]. Thus we have shown (2) Every bornological space is the locally convex hull E[X] = Y,EB B of normed spaces EB. If further, E[X] is sequentially complete, E[X] is the locally convex hull of{B)-spaces. The last assertion follows from § 20,11. (2). By § 19, 3., Yj^b can also be interpreted as a topological inductive limit. B From (2) there follows an important result, due to Mackey [5], which generalises 1.(3). A linear mapping A from a locally convex space E into a locally convex space F is said to be locally bounded if it maps the bounded sets of E into bounded sets of F. (3) A locally convex space E[X] is bornological if and only if every locally hounded map from E[X] into any locally convex space F[_X~\ is continuous. Proof, a) Necessity. Suppose that E[X] is bornological, so that it is equal to £EB. If ^ *s locally bounded, then the restriction of A to EB B is a mapping which sends the unit ball B of EB into a bounded set in FIX]. But then A is continuous, by § 19,1.(7). b) Sufficiency. Putting F[3/] = K, the condition means that every locally bounded linear functional is continuous. Since, further, the identity mapping is a locally bounded mapping from E[X] into E[_(Xk(E')']9 £ coincides with £k(£'). The assertion follows from 1.(3). If we now consider the associated bornological space £pXx], we obtain the following characterisation of locally bounded mappings from general locally convex spaces from (3): (4) A linear mapping from a locally convex space E[X] into a locally convex space F [_X~\ is locally bounded if and only if it is a continuous mapping from E [Z x ] into F [_X~\.
382 § 28. Bornological spaces 3. Local convergence. Sequentially continuous mappings. Criterion 2.(3) for the continuity of linear mappings from bornological spaces can be expressed in another form which is particularly convenient for applications. We introduce the concept of local convergence (which is also called Mackey convergence). A sequence x„ of elements of a locally convex space E[%~\ is said to be locally convergent to x0 if there is a bounded, closed absolutely convex subset B of £[£] such that xn and x0 lie in EB and such that xn converges to x0 with respect to the norm of EB. A sequence which is locally convergent to o is called a local null- sequence. The concept of local convergence clearly depends only upon the dual pair <£',£>. A locally convergent sequence in £[£] is always ^-convergent. (1) a) A sequence xneE[X] is locally convergent to x0 if and only if xn — x0 is a local null-sequence. b) xn is a local null-sequence if and only if there exist positive numbers pn, with p„->oo, such that pnxn is Z-convergent to o. c) // £[£] is a metrizable locally convex space, every X-convergent sequence is locally convergent. Proof, a) If x„ converges to x0 in EB, x„ — x0 converges to o in EB. Conversely, if xn — x0 converges to o in EB, and if Bx is a bounded, closed, absolutely convex set containing #, the x„ and x0, then xn converges to x0 in EBr b) If xn is a local null-sequence, then xnesnB, for some suitable B 1 and some suitable sequence c„->0. But then xn converges to o in EB, and so it converges in £[£]. Conversely, if pnxn is ^-convergent to o, and if B is the closed absolutely convex cover of the pnxn, then xn converges to o in EB. c) follows from a), b) and 1.(5). The problem of giving a precise characterisation of those locally convex spaces in which every convergent sequence is locally convergent is still unanswered. (2) A locally convex space E\1L\ is bornological if and only if every absolutely convex set M which absorbs all the local null-sequences is a ^-neighbourhood ofo. It is enough to show that such a set M absorbs all the bounded sets. Suppose that M absorbs all local null-sequences, but not the bounded x set B. Then for each n2 there is an xneB with -^ $ M. But then xjn is a n
4. Hereditary properties 383 local null-sequence, while on the other hand no scalar multiple of the set of all xjn is contained in M, and this is impossible. A linear mapping from E into F is said to be locally continuous if it maps every local null-sequence into a local null-sequence. (3) If A is a linear mapping from a locally convex space £[£] into a locally convex space F[X~\9 the following properties are equivalent: a) A is locally hounded, b) A is locally continuous, c)A maps every local null- sequence into a sequence which is X'-convergent to o, and d)A maps every local null-sequence into a bounded sequence. Clearly a) implies b), b) implies c), and c) implies d). We must therefore show that d) implies that A is locally bounded. If A were not locally bounded, there would be a bounded set B in E and an absolutely convex neighbourhood V of o in F such that A(B) is contained in no n2 V. Thus there exist points xneB with —y-$V. But then — is a local (x \ n n null-sequence and the AI — J are unbounded, contradicting d). \nj Since continuity of A implies sequential continuity, and since d) follows from this, we have the following result for bornological spaces: (4) A linear mapping A from a bornological space into a locally convex space is continuous if and only if it is sequentially continuous, and if and only if it satisfies one of the conditions a), b), c) or d) of (3). 4. Hereditary properties. The class of bornological spaces is stable under operations similar to those under which the class of barrelled spaces is stable (cf. § 27, 1.). (1) Every locally convex hull of bornological spaces is bornological. Suppose that E[<£~]=YJAa(Fa[%a]\ where each Fa[£a] is bornolo- a gical. If the absolutely convex set U absorbs all the bounded subsets of £[£], then A{a~l)(U) is also absolutely convex, and it absorbs all the bounded subsets of i^pXJ. Thus by hypothesesis A{~l)(U) is a %a- neighbourhood Ua of o in Fa[IJ. But then U=>\~Aa(Ua)9 so that, a by definition of the hull topology, U is a ^-neighbourhood ofo. As a special case, as in § 27, 1. we have (2) Every quotient of a bornological space is bornological. Further, (3) The completion £[£] of a bornological space E[X] is bornological if and only if every locally bounded linear functional on £[£] which vanishes on E is identically zero.
384 § 28. Bornological spaces By § 21,4.(5), 2 is again the Mackey topology, and F=j£jjX])'. The restriction of any locally bounded linear functional on E[X~\ to E is certainly locally bounded on £[2]. By hypothesis it lies in £', so that the assertion follows from 1.(3). The question of whether the topological product of bornological spaces is again bornological leads to some special difficulties. For the time being, we consider the case of countably many factors, and postpone the investigation of the general case until number 8 of this paragraph. (4) The topological product of at most countably many bornological spaces is again bornological. Suppose that the spaces i^pXj, /=1,2,..., are bornological. The 00 topology £ of £[£]= TT i^plj is again the Mackey topology on £, i= 1 by §22,5.(3). By 1.(3), we must also show that every locally bounded linear functional on £[£] is continuous. Suppose that u is locally bounded on £[£]. We assert that there exists an n0 such that uy = 0 for all y = (yi)9 yi^Fh whose first n0 components yt all vanish. If this were not the case, there would be a sequence x(/c) = (x^))g£ with xj-k) = Ofor /=l,...,/c, and ux{k) = k. But the sequence x{k) is bounded in £[£]; this contradicts the local boundedness of u. It is therefore sufficient to establish the result for a product £[£] = ^i[^] x^2p2]- If u is locally bounded on £[£], so are the linear functionals on E defined by m1x = m1(x1,x2) = m(x1,o) and u2x = u(o,x2). But by hypothesis the restrictions of ux to ¥x and u2 to F2 are continuous, so that u1 and u2 are continuous on £, and so also is u = u1+u2. Since every metrizable locally convex space is bornological (1.(4)), we obtain a very extensive class of bornological spaces by repeated application of (1) to (4). For example all (LF)-spaces belong to this class, and so do all spaces of countable degree. Not every closed linear subspace H of a bornological space £fX] is bornological. For example the subspace H2 of the bornological space co(p®(pco considered in §27.2 is not bornological. For, by §27,2., H2 is complete and not barrelled, and so by 1.(2) it is not bornological. Dieudonne [9] has shown that every linear subspace of a bornological space of finite co-dimension is again bornological. The dual of a bornological space need not be bornological, either; indeed there exist (F)-spaces whose duals are not bornological, as is shown by an example in § 31, 7. Amemiya [1] has given an example of a bornological space whose strong bidual is not bornological. 5. The dual, and the topology XCo. In § 21, 6.(4) we showed that the dual E' of every metrizable locally convex space £[£] is £c-complete.
5. The dual, and the topology XCQ 385 By §21,10.(3) the topology Xc on E' coincides with the topology of uniform convergence on the sequences X-convergent to o. But by 3.(l)c) these are the same as the local null-sequences in E. We denote the topology of uniform convergence on all the local null-sequences of a locally convex space E[X~\ by XCo(E); this topology depends only upon the dual pair <£',£>, and does not depend on X. Thus the dual E' of a metrizable locally convex space E is always ^(incomplete. More generally, we now have (1) The dual E' of a bornological space E[X] is Uncomplete. In particular E' is always XC(E) and Xb(E)-complete, and, in the case where E is sequentially complete, E' is also Xk(E)-complete. Proof, a) Suppose that 5 is a XCo-Cauchy filter on E'. It is also a Xs(£)-Cauchy filter, and so it has a limit u0e E* = E \_XS~\, and by § 18,4.(4) this is also the 2Co-limit of g. By 3.(4) it is enough to show that u0 remains bounded on every local null-sequence xn. But there exists a ueE with sup|(w — u0)xn\^e; since u is bounded n on the sequence xn, it follows that u0 is also bounded on the sequence xn. Thus E is XCo-complete; a fortiori it is XC(E)- and Xb(£)-complete, by §18,4.(4). b) Suppose that E[X] is sequentially complete. We must show that Xk(E) is finer than XCo(E), i.e. that the closed absolutely convex covers r~(C) in E[X] of local null-sequences C={xn} are weakly compact. But this follows from § 20, 9.(6). Before proving the converse of (1), we consider some preliminary ideas. If K is a compact subset of E[T], every weaker Hausdorff topology X coincides with X' on K. If K is only precompact with respect to X\ and if X is a weaker Hausdorff on K, then K is still precompact with respect to X, but X need no longer coincide with X' on K. We give an example. Suppose that E[Z~\' and E[X~\ are two normed spaces, and that X is weaker than X'. Suppose that the mapping / from J^pX'] into E[X~] is not one-one (cf. the example of § 18,4.(4)). Suppose that ze/("^(o), z4=o, and that x„ is a sequence of elements in E with X'-limit z. This sequence has X-limit o; the set {o,x1,x2,...} is therefore compact in £[X], but not in £[X']. However, we have (2) Suppose that a second, finer, topology X' is given on the topological vector space £[£]. IfX' has a base ofX-closed neighbourhoods ofo, then X and X' coincide on everyX'-precompact subset M of E. Under the given hypotheses, £[X] can be considered as a linear subspace of E\X\ by §18,4.(4), and so the T-compact closure M of M in E[X'~\ _is a subset of E[X~\. Consequently X induces the same topology on M as does X\ and so X and X' coincide on M. 25 Kothe, Topological Vector Spaces I
386 § 28. Bornological spaces (3) Suppose that two locally convex topologies Xl and X2 are given on the vector space E. IfXl and X2 coincide on an absolutely convex subset M of £, then the uniformities induced on M by Xx and X2 are the same. Proof. A base for the uniformity induced by Xl on M is given by the sets Nv of all (x,y),x,yeM,x — yeU, where U is an absolutely convex ^-neighbourhood of o. Let W be an absolutely convex X2-neigh- W U bourhood of o for which M n-— c M n —, and let Nw be the vicinity corresponding to W. If x\ y'eM, x'— y'eW, then x' — y'elM, x'—y' W U —-—eMn—-c Mn —, so \ha\x' —y'eU, and consequently Nw c Nv. The uniformity induced on M by X2 is therefore finer than the one induced by Xv We now establish the following dual characterisation of bornological spaces : (4) A locally convex space E[X] is bornological if and only if X is the Mackey topology and E' is XCo(E)-complete. One half has been proved in (1). Suppose that E' is XCo-complete. By 1.(3) it is sufficient to show that every locally bounded linear functional u0 on E[X~] belongs to E'. According to Grothendieck's theorem (§21,9.(2)) this is the case if the restrictions of u0 to the sets F(C), where C = {xn} is a local null-sequence in £[2], are weakly continuous. Every local null-sequence C = {xn} is a precompact subset of a suitable EB, where B is a bounded, closed, absolutely convex subset of E[X~\. EB is a normed space; let the norm topology obtained by taking B as unit ball be XB. By §20,6.(2), |"~(C) is also XB-precompact in EB. If we apply (2) to EB and to the two topologies XB and XS(E), it follows that the topologies XB and XS(E') coincide on \~(C). From this it follows, using (3), that the XB-closure of \~(C) in EB is the same as the Xs(F)-closure in EB. Since \~(C)czkB, and since B is weakly closed in E[X~\, it follows that this is the same as the closure \~(C) in E. Thus \~(C) is also precompact in EB, and XB and XS(E') coincide on |~ (C). The restriction of u0 to EB is bounded, and is therefore XB-continuous, and so the restriction of u0 to \~(C)czEB is XS(E')~ continuous. 6. Boundedly closed spaces. If E[X] is a locally convex space, the linear subspace of (F)* consisting of all the linear functional on E which are bounded on the weakly bounded subsets of E is called the bounded closure E of E. E[X~\ is boundedly closed if E = E. The bounded closure of E[X] clearly depends only upon the dual pair <£',£>, and X can be replaced by any other topology compatible with the dual pair (E\E).
6. Boundedly closed spaces 387 Mackey [4] called a dual pair <£',£> for which E = E a "boundedly closed linear system". As a direct consequence of the definition of the associated borno- logical topology Xx (cf. 2.), we have (1) The bounded closure E of a locally convex space E[X] is the dual of the bornological space E[Xkx(E\\. Xfc(£) can be replaced by the weak topology, or any other topology which defines the same bounded sets. There is a simple connection between bornological spaces and boundedly closed spaces; this is made explicit in the next two results. (2) E[X~\ is boundedly closed if and only if E[Xk(Ej] is bornological This follows from the fact that (E'[Zk])' = E, and from the definition of "boundedly closed" and 1.(3). Interchanging E and E in (2), we get (3) £[£] is bornological if and only if X is the Mackey topology and E \%k(E)~] is boundedly closed. Here again, Xk(E) can be replaced by any other compatible topology. 00 We denote the linear span [j nC of an absolutely convex subset C n = 1 of (£')* by L(C). We now have (cf. Dieudonne [9]) (4) The bounded closure E of a locally convex space E[X] is equal to f]L(T) where T runs through all the barrels of E[%~]9 and T stands for T the weak closure of T in (£')*. Every locally bounded linear functional z on E \_XS(E)] lies in (E)*. By §21,2.(1), the barrels T are the closed absolutely convex Xb(E)- neighbourhoods of o in E. Every weakly bounded subset of E therefore lies in some set T°. Since z is bounded on T°, zenT°° for some suitable w, where T°° is the polar of T° in (£')*. By the theorem of bipolars, T°° = f. It follows from this that zef|L(f). Conversely if zef|L(f), z is bounded on all the sets f° = T°. T (5) // £[£] is sequentially complete, E^>E". A boundedly closed sequentially complete locally convex space is always semi-reflexive. For in these circumstances the weakly bounded sets of E are strongly bounded (Mackey's theorem) and every ueE" is bounded on such a set. It follows that if E = E, then E = E". E can also be characterised as follows: (6) The bounded closure E of a locally convex space E\1L\ is equal to the XC0(E [Xj^-completion of E. It follows from (1) and 5.(1) that E is 2Co(F[2fcx])-complete. Since Xk and %£ have the same bounded sets, the local null-sequences in E are the same for either topology; E is therefore XC()(£'[Xfc])-complete. 25*
388 § 28. Bornological spaces We still have to show that E is 2Co-dense in E. By Grothendieck's theorem it is enough to show that every linear functional which is bounded on the weakly bounded subsets of E' is weakly continuous on the sets |"~(C); here C stands for a local null-sequence {un} in £'[JXfc]. But this can be proved in exactly the same way as the proof of 5.(4); instead of E[X] and E, one considers £'|JXfc] and its dual E. (7) A locally convex space E[X] is boundedly closed if and only if it is XCo(£'[jXfc])-complete. Necessity follows from (6). Conversely if E is XCo(£')-complete, E[Xk(E\\ is bornological, by 5.(4), and so E[X] is boundedly closed by (2). 7. Reflexivity and completeness. We now obtain some further conclusions from the results proved in 5. and 6. (1) // the strong dual E[Xb~\ of a locally convex space £[£] is bornological, the bidual E" is %Co(E'\%b\)-complete, and so it is also strongly complete. It follows from 5.(4) that E" is £Co(F[£fc])-complete. The subsets of E' which are strongly bounded with respect to E are the equicontinuous subsets of E' for the strong topology Zb(E',E") on E". But the sets r~(C), where C={un} is a local null-sequence in E[Xb(Ej], belong to this collection, and these sets form a fundamental collection of XCo-equicontinuous subsets of E. Hence Xb(E,E") is finer than Xco; from this it follows that E" is strongly complete. (2) Suppose that E\X\ is sequentially complete. If the strong dual E[Xb~\ is bornological, then E" = E = E[XCo(E)]. Suppose that F|jXb] is bornological. By §20,11.(8), the sequential completeness of E[X~\ ensures that Xb, Xb and Xk coincide on E. It follows from 6.(1) that E = E", and from 6.(6) that E = E[XCo(E)]. In this case the terminology XCo(E) is unambiguous, since the weakly and strongly bounded subsets of E with respect to E are the same, and so the topologies XCo(F[Xfc]) and XCo(F[Xb]) are the same. If £'[XJ is not bornological, and if <Xb(E) = Zk(E"\ then by 1.(3) there is at least one locally bounded linear functional on £'[Xb] which does not belong to E"\ by 6.(1), E" is a proper subspace of E. If the topology %b(E) on E is strictly coarser than Xk(E"), £'[Xfc(£")] can be bornological. In this case we have E" = E = E[XCo(£')]. Komura [2] has shown that, even if E is an (F)-space, we need not always have Xh(E) = Xk(E"). (3) // the strong dual of a quasi-barrelled space E[X] is bornological, then £[2] c E". If E[X] is quasi-barrelled, X = Xh*(E). This topology has the same equicontinuous sets in E as %b{E,E"). The assertion therefore follows from (1).
8. The Mackey-Ulam theorem 389 A special case of (2) and (3) is given by (4) // E[X] is reflexive and £'[JXb] bornological, then E[X] is complete, boundedly closed, and XC0(E)-complete. We now consider (B)- und (F)-spaces briefly. (5) The bidual of a (B)-space E is the bounded closure of E, and is its Xc(E)-completion. Eff = E = E[ZCo(Ef)] follows directly from (2). But the topology XC0(E) on the dual of the (B)-space E is the same as the topology XC(E) of uniform convergence on the compact subsets of E. The assertion E" = E can also be derived simply from the definition of the bounded closure. From (4) and (5) we obtain the following criterion for reflexivity: (6) A (B)-space E is reflexive if and only if it is Xc(E)-complete. A corresponding result also holds for (F)-spaces, but in order to prove it we need the result that the strong dual of a reflexive (F)-space is bornological; this we obtain in § 29.4. In the example of the (FM)-space H(0) considered in § 27,4., we showed directly that (#(£)))' = #(91) is bornological, since we showed (§27,4.(7)) that the strong topology on H(tyL) is the topology of the locally convex hull of a sequence of (B)-spaces. From (2) and 6.(6) there follows directly (7) The bidual E" of an (F)-space E[X^\ is always contained in E = E[iC0{E')l If further, the strong dual of £[£] is bornological, then E" = E = E[±C0(E')l The converse of this last statement is false (see the remark before (3)). 8. The Mackey-Ulam theorem. We now return to the question of when a topological product of bornological spaces is bornological. First we prove two lemmas. (1) // the topological product E [X] = TT Ea [XJ of bornological a. spaces £a[£a] is not bornological, there exists a discontinuous locally bounded linear functional u on E which vanishes on the linear subspace a Since X is the Mackey topology, by §22,5.(3), there exists a discontinuous locally bounded linear functional u on E, by 1.(3). There can only be finitely many £a <= E on which this does not vanish, for otherwise, as in the proof of 4.(4), there would be a sequence xr-e£a. which is bounded in E, but on which u is unbounded.
390 § 28. Bornological spaces The restriction of u to the product TT £ak of the finitely many £ak on which u does not vanish is, by hypothesis, a continuous linear functional on TT£ak. If we extend this linear functional to a linear functional v on the whole of E which vanishes on the product of the other spaces £a, then v is continuous on £, and u — v has the desired properties. (2) Suppose that the spaces £a[£a] are bornological If TT£apa] a is bornological, so is every subproduct TT Ep[Xp], where the fi run through a subset B of the set A of indices a. For every discontinuous locally bounded linear functional on T\Ep can be extended to a linear functional with the same properties on TT£a, by setting it equal to zero on the product of the other spaces £a. Next we show that the question of whether TT£a is bornological or not depends only on the cardinal d of the collection of spaces £a. (3) A product E[X] = T[£a[Xa] of d bornological spaces £a[Xa] a is bornological if and only if cod is bornological. a) Suppose that a>d is bornological. Let us suppose that E is not bornological. By (1) there is a linear functional u with the properties described there. Let x be an arbitrary element of E with components xae£a. The topological product TT^] of the one-dimensional sub- spaces \_Xp] of Ep defined by the non-zero xp is a subspace of E which, under the induced topology X, is topologically isomorphic to a space cod>, d'^d. The restriction u0 of u to TT[^] is locally bounded and vanishes on © [xj. But a>d> is bornological, by (2), and so u0 is continuous. But, by §22,5.(2), a continuous linear functional on a>d, is identically zero if it vanishes on all the elements of a>d. which have at most finitely many non-zero components. Thus ux = uox = 0 for all xeE, which gives a contradiction. b) Suppose that £[£] = TT£a[Xa] is a product of d bornological a spaces ^Paj+fo), and that E[X] is bornological. As we have just seen, a>d is topologically isomorphic to a space TT [xj <= E, where a xa=ho for each a. A locally bounded linear functional u0 on TT [xj a can be extended to a locally bounded linear functional u on E: for each Ea can be written as a topological product \_xa~\ x Fa, so that TT[xa] has a topological complement in TT £a, and it is sufficient to set u equal a to zero on this complement. Since E is bornological, u is continuous on E, and so u0 is continuous on TT [xa]=cod. We denote the topological product KA by co(A) (cf. § 1, 8.), where K is either the real or the complex field, and A is an index set. co(A) is topologically isomorphic to cod, where d is the cardinal of A. If B cz A,
8. The Mackey-Ulam theorem 391 co(B) is identified with the subspace of co(A) consisting of all vectors (£J with £a = 0 for a$B. We say that a linear functional u defined on co(A) vanishes on B cz A if its restriction to co(B) vanishes. We now prove a stronger form of (1) for co(A). (4) If (o(A) is not bornological, there is a discontinuous locally bounded linear functional on co(A) with the following properties: a) u vanishes on all the finite subsets of A, and b) if A is decomposed into two disjoint sets Ax and A2, then u vanishes on just one of the two sets. By (1), there is a u which satisfies a). We assert that there is a decomposition of A into finitely many pairwise disjoint subsets A, such that the restrictions of u to the corresponding spaces co(Af) also have property b). Let us suppose that this is not so. There is then a decomposition of A into two sets Bx and B2 on which u does not vanish. One of the two sets must be decomposable into two sets on which u does not vanish. Repeating this procedure, we obtain a sequence of pairwise disjoint subsets A, of A on which u does not vanish. But using a sequence XiEoiAi) for which ux{^0 we can again construct a bounded sequence in co(A) on which u is not bounded. One of the sets A, has the same cardinal as A; co(Af) is topologically isomorphic to co(A), and the restriction of u to co(Af) has properties a) and b). (5) // <jod is bornological, so is co2d. Let us suppose that co(A) is not bornological, where A has cardinal 2d, and let u be a locally bounded linear functional on co(A) with properties a) and b) of (4). The elements of A can be written as vectors (x = ((xp), where the indices j? form a set B with cardinal d, and where <xp takes the values Oor 1. Each jS determines a decomposition of A into two sets, one consisting of those a for which a^ = 0 and the other those a for which oip=l. By b), u vanishes on exactly one of these two sets, which we shall denote by A^. Let A'p denote its complement. We form the set M = yA^. We can also express M as the union p of at most d pairwise disjoint subsets Yye Ay, on each of which u vanishes. We assert that u also vanishes on M. Suppose that x = (xy)e(o(M), xyEco(ry). The topological product TT[xJ of the one-dimensional spaces corresponding to the non-zero xd is then bornological, by hypothesis. Since u vanishes on the vectors xr u therefore vanishes on each x, and so it vanishes on M.
392 § 29. (F)- and (DF)-spaces By b), u does not vanish on the complement f]A'p of M. But, by p the construction of the A'p, the set (]A]j consists of just one element, p and so we have obtained a contradiction to a). A cardinal d is said to be strongly inaccessible if a) d>K0, b) every sum £dy of d'<d cardinals dy<d is a cardinal less than d, y and c) if f <d, 2f <d. It is in fact not known if any strongly inaccessible cardinals exist. If so, there is a smallest one. We now come to the Mackey-Ulam theorem: (6) The topological product of d bornological spaces is bornological, if d is smaller than the smallest strongly inaccessible cardinal. By (3) it is sufficient to prove this for a>d. If there is a non-bornological co^ there is a smallest cardinal d0 for which a>do is non-bornological. This cardinal has properties a) and c), by 4.(4), and by (5). Since cod- is bornological for d'<d0, the topological product of d spaces codv, with dy<d0, is also bornological, by (3), and so d0 satisfies b) as well. Mackey [3] showed that the question of when TT Ea is bornological is equivalent to a problem in measure theory which Ulam [1] had considered. Theorem (5) is due to Ulam. It is not known if (6) is true for all cardinals d. § 29. (F)- and (DF)-spaces 1. Fundamental sequences of bounded sets. Metrizability. An (F)-space is barrelled and bornological. As a result, theorems about such spaces hold for (F)-spaces. Besides this, we have considered (F)-spaces in detail, within the framework of the general theory. Up to now, however, our knowledge of the properties of the strong dual of an (F)-space has been very slight. In this paragraph we shall develop some of the recent results in the theory of (F)-spaces and their duals, which originated with Grothendieck [2], [10] and with Donoghue and K. T. Smith [1]. We say that a locally convex space E[X] possesses a fundamental sequence of bounded sets if there exists a sequence ^ cz J52 c • • • of bounded sets in E[X] such that every bounded set B is contained in some Bk. In a (B)-space such a fundamental sequence always exists. We now investigate whether this is also possible in (F)-spaces which are not (B)-spaces.
1. Fundamental sequences of bounded sets. Metrizability 393 (1) Suppose that the locally convex space E[X~\ has a fundamental sequence Bx cz B2 <= • • • of bounded sets. If none of the sets Bk absorbs all the other sets Bh there is a countable subset M whose closure is strictly larger than the set obtained by taking all the limits of the Cauchy sequences in M. Proof. We can suppose that the sets Bk are absolutely convex, and that no Bk absorbs the next set Bk+l. Let xn be a sequence of non-zero terms in Bl which converges to o. For each (n, k) let us choose a znke—Bn k with znk((k-\-\)Bn_l. Let M be the set of all xn + znJc, n,k= 1,2,.... For fixed n, xn is obtained as a limit as /c-> oo. xn does not belong to M, so that o is a limit of limits of Cauchy sequences in M. But o is not itself limit of a sequence in M. For on the one hand such a sequence would have to contain elements xn + z„tk with unbounded n; on the other hand the terms would all have to belong to a fixed set Bm, and this would mean that znkeB1+Bmcz2Bm, so that n^m. (2) // £[£] is metrizable and locally convex, and if E[X] has a fundamental sequence of bounded sets, then E[X] is normable. For, by (1), E[X~\ is only metrizable if there is a bounded set B which absorbs all the other bounded sets. But then E \lLh(E)\ is normable with unit ball B\ so that E[Xb*(E)~] is normable, with unit ball B°°. But Z = Zh,(E\ by §21, 5.(3). As a special case, we have (3) An (F)-space is a (B)-space if and only if it has a fundamental sequence of bounded sets. A further consequence of (1) is (4) A locally convex space E [X] with a fundamental sequence of bounded sets is metrizable if and only if its strong dual is normable. The next result deals with the bounded subsets of a metrizable locally convex space: (5) // £[£] is locally convex and metrizable, and if Bn is a sequence of bounded subsets of E, then there always exist positive scalars pn such OO that [j pnBn is also bounded. n= 1 If U1zdU2^>'" is a base of absolutely convex neighbourhoods of 00 o in E[X], and if pnBn c Un, then [j pnBn^ Um for each m, so that oo n = m M pnBn is bounded. n=1 As far as the strongly bounded subsets of the dual of a metrizable locally convex space are concerned, we have
394 § 29. (F)- and (DF)-spaces (6) Suppose that E[Z] is locally convex and metrizable. If Ul => U2 => •" is a base of neighbourhoods ofo in E[Z], the sets U°nform a fundamental sequence of strongly bounded subsets ofE'. For by § 21, 5.(3), every strongly bounded subset M of E is contained in an equicontinuous set U°, where U is a neighbourhood of o, and, since U => C/k for some suitable fc, it follows that M cz [/° <= £/£. (7) // £[£] is locally convex and metrizable, E[_Zb~] is metrizable if and only if E[Z] is normable. For if E [Xb~\ is metrizable, the bidual E" is normable in the strong topology Zb(E',E"), by (4), and so therefore is E[Z], since Z = Zb*(E) (§21,5.(3)), and since the topology Zb(E,E") induces the topology Zb*(E) on E, by §23, 4.(4). (8) Suppose that E[Z] is locally convex and metrizable, but not normable. Then there is a fundamental sequence Bx cz B2 cz •■• of strongly bounded absolutely convex weakly closed subsets of E, none of which absorbs the next, and such that each EBn is a proper subspace of E'Bn+l. 00 Here EBn denotes the linear span of Bn, and E= \J EBn. n= 1 Proof. It follows from (6) and § 21, 5.(3) that there exists a sequence (Bn) such that no Bn absorbs the next set Bn+l. By § 21, 6.(4) and § 20,11. (2), EBn is a (B)-space when Bn is taken as unit ball. If EBn were equal to EBn + i, EBn would also be a (B)-space under the strictly coarser norm topology of EBn + x, and this contradicts § 15,12.(7). (9) An (F)-space E is a (B)-space if and only if E contains a bounded absorbent set. The condition is clearly necessary. Conversely if E contains a bounded absorbent set, E also contains a bounded barrel B. Since E is barrelled, B is a bounded neighbourhood ofo in E, and E is therefore a (B)-space, by §15,10.(4). 2. The bidual. Suppose that U1=>U2=>'" is a base of absolutely convex neighbourhoods of o for the metrizable locally convex space £[£]. Then the sequence t/?° => U°2° => ••• of polars in E" of the sets U° forms a base of neighbourhoods ofo for the natural topology Zn(E) on the bidual E" (cf. §23,4.). Since £[£] is metrizable, Zn{E) is the same as the strong topology Zb(E,E") onE". (1) Suppose that E[_Z~\ is locally convex and metrizable, with strong 00 dual E[_Zb(E)~\ and strong bidual E"\Zn(E)~\. If the union M = [j Mn M= 1 of countably many Zb(E)-equicontinuous subsets Mn of E" is ZS(E)- bounded in E", then M is also Zb(E)-equicontinuous.
2. The bidual 395 Proof. We can suppose that Mn a Mn+1, and that the sets Mn and M are absolutely convex. The equicontinuity of Mn means that Mn a J3°°, where Bn is an absolutely convex bounded subset of £[£], and B°n° is the polar of B°n c E in E". The set M is weakly bounded, by hypothesis; since £'[£&] is complete, it is bounded in £"[£„], by the Banach- Mackey theorem. Thus for each ^-neighbourhood U°k° there is a ck>0 with M c ck U°k°. Further, for each k and n, there exists an ank>o with BnaankUk. Let bk = max(ck,ank). Then f| ^^fc^ 0 ^^ / oo \o / oo \o so that [ f] bkUk\ cz( f] ankUk) cz B°n. It follows from this that \n + 1 / \n + 1 / M^(c)bkuY \n+l J 1 On the other hand it follows from M„c M a bk U°k° that M„°D-[/fc°, " f1 \ /A V bfe so that MJd [-_[/£= (I fct t/t ; this last equation follows k=i\h J \k=i J / " Y from §20,6.(5) and §20,8.(10). Consequently 2M°n => I f] bkUk 1 /oo \o / n \ o \fc = 1 / + I P) bfe C/k I . But now I f] bk Uk) is the polar of a neighbourhood of o, and so it is weakly compact in £'. Thus I f) bkUk\ +1 [) bk Uk J is weakly closed, by §15,6.(10). But it follows from this that / oo \o / n \o / oo \o U\bkUk\ =[C\bkUkJ + f f) h Uk) , and consequently 2M^(f)bkUk)\ \k=1 J / oo \oo Zoo \oo Thus for each n we have iM„c (l^^) ,andso^Mczl ()bnL/fe) . 00 \fc=i / \fc=i / But since f] t^ Uk is a bounded subset of £, our assertion is proved. k= 1 (1) can also be expressed as a property of £'[£,,]: (2) Suppose that £[£] is locally convex and metrizable, with strong dual £'[£J. // Vn is a sequence of absolutely convex strongly closed neighbourhoods ofo in £'[£&] whose intersection V is absorbent, then V is also a neighbourhood ofo. Proof. If we put Mn=V°n, Mn is a ^-equicontinuous subset of E". / oo \o oo V° = i f]Vn) id f]V°n = M. Since V is absorbent, V° is weakly \n=\ J n=l bounded in £", so that M is weakly bounded. If we now apply (1), we obtain (2). Similarly it can be shown that (1) follows from (2).
396 § 29. (F)- and (DF)-spaces It follows from (1) that every countable bounded subset of £"[!„] is £b(£)-equicontinuous, and so is relatively £s(£')-compact. In particular every weak Cauchy sequence and every £„-Cauchy sequence is a set of this kind; such sequences therefore always have a limit in E". Since £"[£„] is metrizable, we therefore have (3) If E\1L\ is locally convex and metrizable, its bidual E"[Xn(E')'] is a weakly sequentially complete (F)-space. In particular the strong bidual of an (F)-space is again an (F)-space. Further, we have (4) If E\X\ is a non-reflexive (F)-space, the strong bidual E"[Xn(E')~\ is again a non-reflexive (F)-space. Proof. Let us suppose that E" is reflexive. By § 23, 2.(5), the topology %k(E") on E is finer than Zb(E). But E[Zb(E)] is complete, so that F[2fe(£")] is also complete. If we apply §23,5.(6) to E[Xk(E% it follows that E[Xk(E")] is reflexive. Consequently E[%h(Ey\ is semi- reflexive. Another application of § 23, 5.(6) shows that £[£] is reflexive, which gives a contradiction. From (4) and § 23, 5.(7), we obtain the following result, analogous to the one for (B)-spaces: (5) If E is a non-reflexive (F)-space, the iterated strong duals are all non-reflexive, and in each of the sequences E c E" c E" c ••• and Fc E" c ■•■ each space is a proper subspace of its successor. 3. (DF)-spaces. In 1.(6) and 2.(1) we obtained two special properties of the strong dual of a metrizable locally convex space; these were used by Grothendieck to define a new class of locally convex spaces. A locally convex space E\X\ is said to be a (DF)-space if a) it has a fundamental sequence of bounded sets, and b) every strongly bounded subset M of E which is the union of countably many equicontinuous sets is also equicontinuous. As in 2.(2), it is easy to see that b) can be replaced by the dual property b'): if Un is a sequence of closed absolutely convex neighbourhoods ofo 00 in £[£], and if [/=(](/„ absorbs every bounded set, then U is a n= 1 ^-neighbourhood ofo. A quasi-barrelled locally convex space with a fundamental sequence of bounded sets is always a (DF)-space. Thus every normed space is a (DF)-space. By 2.(1), the strong dual of a metrizable space is a complete (DF)- space.
3. (DF)-spaces 397 As in 2, it follows from a) and b) that (1) If E[%] is a (DF)-space, £'[£J is an (F)-space. The next result, about the neighbourhoods of o in a (DF)-space, is due to Grothendieck [10]: (2) Suppose that £[£] isa(DF)-space,andthat (Bn) is a fundamental sequence of closed, absolutely convex bounded subsets of E. An absolutely convex subset WofE is a %-neighbourhood ofo if and only if Wn Bn is a X-neighbourhood ofo in Bn, for each n. Proof. We have to show that the condition is sufficient. For this, we construct a sequence ocn of positive numbers and a sequence Un of closed absolutely convex neighbourhoods ofo, such that (3) anflnc:£W, (4) anBnczUk, (5) UHnBnczW for all n and all k. The proof of (2) follows from this construction; for if 00 we put U= (°) Un, then by (4) U absorbs every bounded set, and so n= 1 by b') U is a ^-neighbourhood ofo; finally it follows from UnBna W that Un( Q B„J=C/c W. Suppose that an and Un have been defined for nrgm in such a way that (3), (4) and (5) are satisfied for n, k^m. By hypothesis, there is a neighbourhood U of o with U nBm+1 c W. We choose ocm+1 in such a way that am+1 Bm+1 c ±U and <*m+1Bm+lcz%Bm+l. But then (xm+lBm+1 <=i(#m+i nU)a^W, so that (3) is satisfied for m+1. Further am+! can clearly be chosen small enough for (4) to hold for n = m +1 and /crgm. m+ 1 We set J5(m+1)= |~~ a,-£f- If we can find an absolutely convex neigh- i= 1 bourhood Kofo such that Um+1 = J3(m+1) + V satisfies condition (5) for m+1, then since anBn a B(m+1) c C/m+1, (4) is also satisfied for n^m + 1 and k = m+l. Since t/m+1 c 2J3(m+1) + 2 J/ it is sufficient to show that (2B(m+1) + 2K)nJ3m+1 c W. If we set M = Bm+ln(E~W), we must show that (2J3(m + 1) + 2 K)nM is empty for a suitable choice of V, i.e. that 2Fn(M + 2J3(m+1)) is empty. This means that we must show that N = M + 2Bim+1) cannot haveo as a closure point. We prove this in the following way. Since B(m+1) c \W, we have ^W+2Bim+1)^W. Since WnM is empty, (±W+2B{m+1))nM is also empty, and so Nn^W is empty. The set 3N is bounded, and so it is contained in some Bk; also WnBk is a neighbourhood ofo in Bk.
398 § 29. (F)- and (DF)-spaces Since (3N)nW is empty, o is not a closure point of 3 N, and so it is not a closure point of N. Let us state the result for the special case of the strong dual of a metrizable locally convex space explicitly: (6) Suppose that E[%] is locally convex and metrizable, and that t/j => l/2 => '" is a base of absolutely convex neighbourhoods of o. An absolutely convex subset W of E is a strong neighbourhood of o in E if and only if Wn U°n is a strong neighbourhood ofo in t/°, for each n. As a simple consequence of (2), we have (7) A linear mapping A from a (DF)-space E into a locally convex space F is continuous if and only if its restrictions to the sets Bn of a fundamental sequence of bounded sets are continuous. For if A is continuous on the sets Bn, and if Kis an absolutely convex neighbourhood ofo in F, A{~l)(V)nBn is a neighbourhood ofo in Bn; by (2), the absolutely convex set A{~X)(V) is thus a neighbourhood of o in E. The following result, which is also due to Grothendieck, relates the topology of a (DF)-space to the topology Zb*(E): (8) If M is a separable subset of a (DF)-space £[£], the topologies *% and Zb*(E) coincide on M. Proof. Xh* is finer than X We must therefore show that given a closed absolutely convex ^-neighbourhood V of o there is always an open ^-neighbourhood U of o with Mn[/c V. This is equivalent to the assertion that (Un(E~V))nM is empty. Since Un(E~V) is open, it is sufficient to show that if x( is a dense sequence in M, no element of the sequence lies in U n(£~ V). In other words, we must show that there is a U which contains none of the x„ which lie in E~V. We denote the sequence xik by x1,x2,... again. There is clearly nothing to prove, unless this sequence does not terminate. For the proof, we construct sequences a„>0 and Un (closed absolutely convex ^-neighbourhoods ofo) such that (9) «nBnczUk, (10) anBn^V, (11) xniUn for all n and k; the sets Bn are a fixed fundamental sequence of bounded closed absolutely convex subsets of E. Suppose that these have been found for k,n^m. It is clearly possible to choose am + 1 in such a way that (10) holds for m + 1 and (9) holds for n = m + l and /crgm. tn + l We set B{m+ ]= [~ anBn. We now have to choose a closed absolutely convex Um + l => B{m + l) in such a way that xm + l$Um + l. But this can be done, since J5(m + 1) c V and xw + 1 $K (cf. § 15,6.(9)).
4. Bornological (DF)-spaces 399 If we put U = f] Un9 then because of (9) we obtain an absolutely convex set which absorbs all the bounded sets, and which is therefore a ^-neighbourhood, by b'). Because of (11), U contains none of the xne£~K Using (8), sufficient conditions can now be given for a (DF)-space to be quasi-barrelled, i. e. for % and %b*(E') to coincide. (12) a) Every separable (DF)-space is quasi-barrelled. b) // the bounded subsets of a (DF)-space E[X] are metrizable under X, £[£] is quasi-barrelled. c) An (F)-space is distinguished if the bounded subsets of the strong dual are metrizable. Proof, a) follows directly from (8). By (2), %b*(E') is the same as % if %b*(E') and X coincides on any bounded set B. If % is metrizable on J5, % is determined by the collection of sequentially closed subsets of B. But the ^-convergent sequences are the same as the ^-convergent sequences, by (8). Since every ^-closed set is £b*-closed, the ^-closed and £b*-closed sets of B are the same, so that X and Xh* are the same on B. Thus b) is proved. By §23,7.(1), an (F)-space E is distinguished if and only if FpIJ is barrelled. Since E' is complete, c) follows from b). 4. Bornological (DF)-spaces. Suppose that E\%\ is a metrizable locally convex space, that E'[Zh(E)] is its strong dual, and that £"[£„] is its strong bidual. We now investigate the bornological space associated with F[2J. (1) Let Bn be a fundamental sequence of absolutely convex bounded closed subsets of F[3J. The algebraic hull Va of any set V of the form 00 V = r~ otnBn, a„^0, is always Xh-closed. n= 1 Thus Ef[%b] has a base of ^-closed neighbourhoods of o. k °° Proof. Let Vk= [~ anJ5n, so that V= U Vk. Since the sets Bk n=1 fc=i are £s(£)-compact (cf. §21,5.(3)), the sets Vk are also £s(£)-compact, by §20,6.(5), and so they are strongly closed. Suppose that u does not belong to Va. Then by § 16,4.(4) there exists a /?>1 with u$fiV. Since u$fiVk, there exists a zkeVk c E" with zku = fi. The sequence zk is bounded in £"[£„], and so by 2.(1) it is an equicontinuous, and therefore relatively £s(£')-compact, subset of E". If z0eE" is weakly adher- 00 ent to the sequence zk, z0u = (3, and z0e f] Vk°= V°. Thus u does not k= 1
400 § 29. (F)- and (DF)-spaces belong to V°\ and the ^-closure Voc of V is equal to Va. Finally the sets Va form a base of neighbourhoods of o for Zbx, by §28,2. (1) now enables us to determine the topology Zhx on £'. (2) Suppose that £[£] is metrizable and locally convex. The bornological topology Xhx associated with the strong topology Zb(E) on E' is equal to Hb{E"). E'\%b(E")\ is thus always a complete bornological (DF)-space. Proof. By §21,2.(1), the closed absolutely convex absorbent subsets of F[£b(£)] form a base of £b(£")-neighbourhoods of o in £'[2b(£")]. But, by (1) and the definition of bornological space (§28,1.), these sets also form a base of ^-neighbourhoods of o in E'. The completeness of E'[Xb(E")'] follows from the ^(^-completeness of £', using (1) and § 18, 4.(4). We now use (2) to give a criterion for the strong dual of an (F)-space to be bornological, i. e. for it to be an (LB)-space. (3) Suppose that E[X] is an (F)-space. The following are equivalent : a) E[X] is distinguished; b) £'[JXb(£)] is bornological; c) £'[£&] is barrelled, or quasi-barrelled. Proof. Since F[IJ is complete, a) and c) are equivalent, by §23,7.(1). Since every bornological space is quasi-barrelled (§28,1.), c) follows from b). On the other hand, b) follows from c), by (2). It follows from 3.(12) that if E,[Xh] is separable or if the bounded sets of £'[£J are metrizable, then £'[£J is bornological. Further we have (cf. § 28, 7.): (4) // E is a reflexive (F)-space, its strong dual is bornological. An (F)-space E is reflexive if and only if it is boundedly closed, and if and only if it is %Co(E')-complete. Proof. If E[X~\ is reflexive, £'[£&] is barrelled, and so by (3) it is bornological. It follows from §28,7.(4) that a reflexive space E is boundedly closed and £Co-complete. On the other hand if E = E = E[XC0(E% then E = E\ by §28,7.(7). In §31,7. we shall give an example of an (F)-space which is not distinguished. Thus there are (F)-spaces E with the property that not every bounded set M of E" is contained in the bipolar J5°° of a bounded subset B of E. If however M is weakly separable, it follows from the definition of (DF)-space that there always exists such a B, with M c J5°°.
5. Hereditary properties of (DF)-spaces 401 (5) // £[£] is an (F)-space, and if E is its bounded closure, then every Xs(E')-bounded subset of E lies in the bipolar Boc in E of a bounded subset B of E". By (2), E'\Xh{E")\ is barrelled, and by § 28, 6.(1) its dual is E; thus the bounded sets of E are the same as the £fo(£")-equicontinuous sets. These are the subsets of sets B°°, where B° is a £fe(£")-neighbourhood of o. (6) Suppose that E[X~\ is locally convex and metrizable. E" = E = E[XCo(E,j] if and only if Xk(E") = Xb(E"), and if and only if every Xs(E')-bounded subset of E" is relatively Xs(E')-compact. This follows from (2), since £ = (£'p;])' and E" = {E'\Xk{E")\)'. In 3. we gave criteria for the topology X of a (DF)-space to be equal to Xh*(E'), and in the present number we considered situations in which X is bornological. The following simple example shows that in general X can be coarser than the Mackey topology. Suppose that £ is a reflexive (B)-space which is not separable. We give E the topology of uniform convergence on the separable bounded subsets of the strong dual E'. Then E[X~\ is a semi-reflexive (DF)-space, and X is strictly coarser than the topology Xk(E') = Xb(E'). 5. Hereditary properties of (DF )-spaces. (1) If E[X~\ is a (DF)-space and if H is a closed linear subspace, then (E/H) [£] is also a (DF)-space, and the two strong topologies Xb(E) and Xb(E/H) on H1 = (E/H)' are the same. First we prove the second part of the theorem. The topology Xb(E) on H1 c E' is coarser than Xh(E/H), so that the identity mapping / from H1[Xb(E/H)'] to H1[Xb(E)'] is continuous. We must show that its inverse is also continuous; then / is a topological isomorphism. H1\Xh(E)~] is metrizable, as it is a subspace of F, and so it is bornological; by §28, 3.(4) it is sufficient to show that every sequence un in H1 which is £fo(£)-convergent to o is bounded in H1[Xb(E/H)~]. Since E[X~\ is a (DF)-space, though, the set {un} is £-equicontinuous in H1. As such, it is relatively £s(£)-compact, and so it is also relatively Xs(E/H)-com- pact; consequently it is ^(£/H)-bounded. Thus we have shown that on H\Xb(E) = Xb(E/H). From this it now follows that every bounded set in (E/H)[X] is contained in the closure of the canonical image K(B) of a bounded set B of £[£]. Thus (E/H)[X] also has a fundamental sequence of bounded sets. Finally, condition b) of the definition of (DF)-space is satisfied by E/H, since it holds for £. A closed linear subspace of a (DF)-space need not be a (DF)-space (cf. the counterexample in § 31, 5.). (2) // the linear subspace H[X] of a locally convex space E[X] is a (DF)-space, the strong dual of H[X~\ is topologically isomorphic to E'[%(£)]/H\ 26 Kothe, Topological Vector Spaces I
402 § 29. (F)- and (DF)-spaces We must show that the algebraic isomorphism of H'[Xb(H)] onto E'\Zb(E)\IHL is a topological isomorphism. Since ^(H) is coarser than Zb(E) on H\ it is sufficient to show that the isomorphism is continuous. Since H'[%b{H)~\ is metrizable, and so is bornological, we need only show that any sequence un which converges to o in H'\%h(H)~\ is bounded in the topology %b(E). Once again, the un form a £-equi- continuous subset of H', and this set is the image of a £-equicontinuous subset of E', by §22,1.(1). This set is £b(£)-bounded, and so also is its image {un}. (3) a) A (DF)-space £[£] is complete if and only if it is quasi- complete. b) The completion of a (DF)-space is again a (DF)-space. c) Every semi-reflexive (DF)-space is complete. Proof, a) We apply (2) to E and to its completion E. The two strong topologies Zb(E) and Zb(E) coincide on E' = E'. Thus every bounded set in E lies in the closure of a bounded set in E. If E is quasi-complete, it follows that E = E. b) If Bn is a fundamental sequence of bounded sets in £, the closures Bn in E form a fundamental sequence of bounded sets in E. Condition b') for a (DF)-space is satisfied by £, if it is satisfied by E. c) Every semi-reflexive space is quasi-complete, so that c) follows from a). (4) The locally convex hull E[<X] = YJAn(E„[%„]) of a sequence of (DF)-spaces £„[£„] is again a (DF)-space. Every bounded subset of E lies in the closed absolutely convex cover of finitely many An(Bn), where Bn is bounded in £„[£„]. The strong dual E'[%b(E)] is the locally convex kernel of the spaces Ai-l\E'n\%b{En)-\). 00 First we show that the locally convex direct sum F = © £„[£„] of (DF)-spaces is again a (DF)-space. As J5|n) runs through a fundamental sequence of bounded sets in N EnV^n]) tne sets © Bin\ taken in a suitable order, form a fundamental n= 1 sequence in F, by § 18, 5.(4). The fact that F satisfies condition b') for a (DF)-space follows from the fact that the spaces £„[£„] satisfy b'), and from the definition of the locally convex direct sum topology. Since ^An(£n[Xn]) is topologically isomorphic to a quotient F/H, the first assertion of (4) follows from (1). Likewise it follows from (1) that every bounded subset of F/H lies in the closure of the canonical image of a bounded set in F, and so
6. Further results, and open questions 403 it lies in a set of the form f~ 4f(jBf). The last part of (4) follows directly i= 1 from this, when we apply § 22, 7.(5). In particular (4) says that every topological inductive limit £[£] of an increasing sequence fjpj] c: £2[^-2] c: ••• of normed spaces is a (DF)-space. By §28,4.(1), E[X] is then bornological. We remark that £[£] need not be complete, even when the spaces En[Zn] are (B)-spaces (cf. § 31, 6.). (5) A locally convex space E[X] is a bornological (DF)-space if and only if it is the topological inductive limit of an increasing sequence of normed spaces. The simple proof that the condition is also necessary is left to the reader. (6) The locally convex hull of a sequence of semi-reflexive ( respectively reflexive) (DF)-spaces En is again a semi-reflexive ( respectively reflexive) (DF)-space. If the spaces £„[£„] are semi-reflexive (DF)-spaces, every bounded set £„(=£„[£„] is relatively Xs(£^)-compact. But then the absolutely convex cover of finitely many An(Bn) is relatively £s(£')-compact, so that YjAniEnl^n]) is semi-reflexive, by (4). If the spaces £„[£„] are reflexive, they are barrelled. By §27,1.(3), Y^^Jfin[£„]) is also barrelled. The assertion now follows from §23,5.(1). 6. Further results, and open questions. The properties of linear mappings between (F)- and (DF)-spaces will be investigated in the second volume, and important special classes of these spaces will also be considered there. For the present, we make a few additional remarks. If E is a normed space with completion £, every bounded subset of E is contained in the completion of a bounded subset of E. The question of whether this is also true for metrizable locally convex spaces was settled quite recently. (1) // £[^] is separable, metrizable and locally convex, every bounded subset of the (F)-space E[*X] is contained in the completion of a bounded subset of E. As Grothendieck [10] showed, this is a simple consequence of the theorems of number 2. E a E", by 2.(3). A bounded subset B of E is thus a separable bounded subset of E". A countable dense subset of B is £b(£)-equicontinuous, by 2.(1); thus J5, being contained in its closure, is £b(£)-equicontinuous. Thus B c M°° n£, where M°° is the polar of M° in E", and M is bounded and absolutely convex in E[%]. But M°° n E is the completion of M in £, by the theorem of bipolars. 26*
404 § 29. (F)- and (DF)-spaces Amemiya [1] has given the following example, which shows that (1) need no longer be true in the non-separable case. Let £ip] be a real (F)-space which is not a (B)-space, and let Ba, txe A, be a fundamental system of absolutely convex bounded subsets of Ex, which must be uncountable, by 1.(3). By 1.(9), for each a there is a discontinuous linear functional uaeEf which vanishes on the elements of Ba. Let p„(x) be a sequence of semi-norms on Ex which define the topology X. We denote by E = llA(E1) the space of all x = (xjy xaeEu for which all the sums q„{x) = YJPn(xa) are finite. E is an (F)-space under the topology defined by the a seminorms qn(x) on E. We denote by coA the space of all real vectors i = ({J, ae A. The equation A(x) = {uaxa) defines a linear mapping from E into coA. Let F = l\ denote the subspace of all xea)A with £|<!:a|<oo. F is a (B)-space under the norm £|£J. a a. Finally let E0 be the subspace of E which is mapped into F by A, E0 is dense in £, since all the terms x = (xj with only finitely many non-zero xa belong to E0. For the same reason A(E0) is dense in F, where A denotes the restriction of A to E0. Now let B be a bounded subset of E0. We assert that A(B) contains no neighbourhood of o in F. If qn(x)^Mn for all xeBy then pn(xa)^M„ for each a; all the components xa of the elements x of 5 thus lie in a fixed bounded set Bao. But then waoxao = 0, for all xeB; thus the a0-th co-ordinate vanishes for all elements of A(B). It follows from this that A(B) can contain no scalar multiple of the unit ball of F. We now consider the graph G(A) of the mapping A from E0 into^F as a linear subspace of the (F)-space Ex F. Since E0 is dense in £, since A(E0) is dense in F, and since, because the functionals ua are discontinuous, /J(-1)(o) is dense in F0, G(A) is dense in Ex F. We assert that no bounded subset M of G(A) is dense in a bounded subset of Ex F of the form Mx x K, where K is the unit ball of F. The set M is contained in a set of the form B x A(B), where B is bounded in E0. But we have previously shown that A(B) is not dense in X, so that M is not dense in M1xK; this establishes the counterexample. Komura [2] has given an example, assuming the continuum hypothesis to be true, of an (F)-space which is not separable, but all of whose bounded sets are separable; earlier Dieudonne [14] had given an example, which again used the continuum hypothesis, of a metrizable locally convex space with the same properties. The question, raised by Grothendieck [10], of whether the topologies £fe(£) and %k(E") on the strong dual E' of an (F)-space are always the same (cf. § 28,7.) was also answered in the negative by Komura [2]. The properties "quasi-barrelled" and "bornological" are equivalent for the strong duals of (F)-spaces, by 4.(3). Komura [1] has shown however that there exist barrelled (DF)-spaces which are not bornological. Amemiya [1] has given an example of a reflexive (F)-space, in which no bounded set is total: let R be the set of monotonically increasing
1. The a-dual. Examples 405 sequences v) = (rjn), rjn>0. Let us consider the vector space E of all real functions f(v)) on R which satisfy pt(f)=( £ |/(t))|2^/)T<oo, for /= 1,2,.... Since each one of these conditions defines a (B)-space which is norm-isomorphic to a space /J, the space £, equipped with the norms Pi(f), i= 1,2,..., is a reflexive (F)-space, as it is the intersection of reflexive (B)-spaces. It is sufficient to show that the elements / of a bounded set 5c£ all vanish at some fixed t)0. If suppt(f) = Mh and if £ = (£,) is chosen /eB_ in K in such a way that limMf ^ 1=0, then f(x) must be zero for i-*oo each feB, since we have the inequalities \f(x)\2^t^Mf, for /=1,2,.... (2) // E is a reflexive (F)-space in which no bounded set is total, the strong dual possesses bounded subsets which are not metrizable under the strong topology. Let Un, n=l,2,..., be a base of neighbourhoods ofo in E. It is sufficient to show that not all the sets U° are metrizable. If they were all metrizable, there would be bounded sets Bnk in E such that the sets B°nknU° form a strong base of neighbourhoods ofo in U°. By 1.(5), there is a single bounded set B in E which absorbs all the sets Bnk. If u is any non-zero element of E\ uel)°n for some n, and there is a Bnk with u$B°nk. In particular, u does not vanish identically on Bnk, and so it does not vanish identically on B. Thus B would be total in £, which contradicts the assumption made about E. This result of Amemiya's answers a question posed by Grothendieck [10]. § 30. Perfect spaces 1. The a-dual. Examples. In this paragraph we consider sequence spaces (which are also called coordinate spaces); these are vector spaces / whose elements are sequences x = (xi) = (x1,x2,-..) of real or complex numbers. The vector space operations are given by the usual operations on the coordinates. We can always consider such a sequence space / as a linear subspace of the space at of all sequences, / c at. A sequence space is said to be n o r m a 1 if whenever it contains x = (xt) it also contains all vectors r) = (yi) with \yt\S \xt\ for /= 1,2,.... For example the space <p of all sequences with only finitely many non-zero coordinates is normal. To each sequence space X we assign another sequence space Xa = kx, its a-dual. Ax is defined to be the set of all sequences u = (ui) for which the
406 § 30. Perfect spaces 00 scalar products ui= ^ utxt converge absolutely, for all xeX. For ; = i example cox = <p and <px = co. It follows directly from the definition that (1) a) If X a fx, then \ix c Xx. fej We always have Xx x = (XX)X =) A. A sequence space / is said to be perfect if Xx x = X. By the remarks above, <p and co are perfect spaces. (2) The oc-dual Xx of an arbitrary sequence space is always perfect. For any sequence space a, we have Xx = a x x x . Xx x is the smallest perfect space containing X. Proof. By (1) a) it follows from Xxx => a that (xxx)x c=/l\ On the other hand, by (1) b), (Xx)x x => Xx, so that Xx = AX x x; X* is perfect. If jj. is perfect and //da, fx = fix x ^> Xx x. But Ax x is perfect, and so it is the smallest perfect space containing X. (3) If X is perfect, X is normal and X => <p. This is immediately obvious for Xx, and so it also holds for Xx x = (Xx)x. Thus every perfect space X satisfies <p c X c co. We now give some simple examples. (4) I1 and /°° are perfect; we have (/1)x = Z00,!/00)" =/1. 00 Since e = (l,l,...)e/°°, it follows that £|u;|<oo for each ue(n*, i=l so that (/°°)x c= I1. On the other hand, clearly I1 c (/oc)x. Likewise it is trivially true that /°° c (/1)x. On the other hand given an unbounded sequence t> =(vt) there is an xel1 for which t> x = ^i;lxf diverges, so that /00=(/1)x. (5) /2 is perfect, and it is the only self'a-dual sequence space. a) The inequality |w„x„|^ |wj2+ |x„|2 implies that ux is absolutely convergent, when both x and u belong to I2. Thus (/2)x => I2. Let us suppose that (/2)x is bigger than I2. Then there is an element o = (u,-) in (/2)x for which £|^-|2 = oo. We can therefore find 0 = n1<n2<"' such that |ull. + 1|2+•••+ |ull. + 1|2 = M?^l. If we put Xj = tt-tVj for 00 ' 1 I All nt-\-l^j^ni+1, we get £ l*/l2 = Z"^' < °°> so tnat * = (x;)ef2- But 7=1 * oo 2 on the other hand £|u,-x,-| g: £ — = oo, which contradicts the as- i = i i sumption that t>e(/2)x. b) Suppose that X = XX. X is therefore perfect and normal. If x = (xt)eX, then x = {Xi)eX = Xx as well, and so we must have xx = £|xf|2<oo.
2. The normal topology of a sequence space 407 From this it follows that X a I2; on the other hand it follows from (1) a) that X = XX z>(l2)x=l2. (6) V is perfect, for 1 <p < oo, and (lp)x = lq, 1— = 1. p q It follows from Holder's inequality that (lp)x =) lq. For a ve(lp)x with 00 Y, \vt\q=oo we proceed in the same way as in (5): we determine sections i= 1 of d satisfying |i;B1+1|«+-+ |t;lli + 1|« = Mf^l, put Xj = |^r for nt + 1 ^j^ni+l, and arrive at a contradiction, as before. 2. The normal topology of a sequence space. It is easy to see how the ideas which have just been introduced can be fitted into the theory of locally convex spaces. If X is a sequence space, and if we further suppose that X contains cp, then X and its a-dual Xx form a dual pair <AX,A>, with bilinear form 00 (u,x} = ux = Yj uixi- The condition X => cp is necessary in order to make sure that condition (D2") of § 10, 3. is satisfied. The construction of Xx therefore enables us to pass from a space X =d cp to a dual pair <Ax,a>, in a unique way. Consequently if X is a sequence space, all those concepts which only depend upon the dual pair, such as the weak topology, the Mackey topology and the strong topology, are also uniquely defined. On the other hand, the way that the a-dual was defined makes it natural to consider the locally convex topology defined by the seminorms 00 (i) ?„(*)= E kl W, uer, i= l on the sequence space X. We call this the normal topology % of the sequence space /. A coordinate vector x = (xt) is said to be positive, and we write x>o, if all the xt are greater than or equal to 0, and if x^o. The sets [/u £ consisting of all xeX with pu(x)^z form a base of ^-neighbourhoods n ofo, as u runs through the positive elements of Xx. For f] UUuE contains the neighbourhood UUiE where u= £ uf. i = l If M is a collection of vectors in co, the set of all t) = (>>;)ecu with |>\.|^|xf|, i=l,2,..., for some x = (xt) in M is called the normal cover
408 § 30. Perfect spaces Mn of M. If M = Mn, M is said to be normal. If a normal space contains M, it contains Mn. If we denote by 9i the class of normal covers {u}" of sets consisting of just one positive element u of Xx, together with their subsets, we clearly have (2) The normal topology X of a sequence space X is the topology X^ of uniform convergence on the collection 9i of normal covers of the positive elements of Xx, together with their subsets. Further, (3) If X=> q>, 9i is the collection of all %-equicontinuous subsets ofXx. It is sufficient to show that £/°? t = {u}", for the collection of £-equi- continuous sets consists of the polars of a base of neighbourhoods of o, together with their subsets. If the z'-th coordinate u{ of u is non-zero, every xtt{ with |xf| = — kl lies in Uul. From this it follows that if veU°tl, then \vt\^ |wf|. If ut = 0, then any multiple of cf lies in UUtl, so that vt must be zero, as well. This proves (3). If X = a), then Xx =q>, and the normal topology on at is the same as both the weak and the strong topologies (with respect to cp), and so it 00 is the same as the topology of the product TT £„, En= K. n= 1 If X = cp, then Xx = co, and the normal topology on cp is the same as the strong topology, and so it is the same as the locally convex direct sum topology on cp. If X = l\ then Xx=l^ by 1.(4), and the normal topology on Z1 is 00 the same as the topology given by the norm ||x||= Y, \xh so tnat ^ ^s the same as the strong topology. £= x In the case where X = lco, so that Xx=ll, the collection of sets 91 does not consist of all the norm-bounded subsets of/1. Indeed, by § 22, 4.(3), 91 is a proper subset of the collection of all £s(/°°)-compact subsets of I1. Thus the normal topology on /°° is strictly coarser than ^(Z1). On the lp spaces, \<p<cc, the normal topology X is also strictly coarser than the Mackey topology, as can easily be seen. If a is a sequence space and if c = (cf) is a vector with non-zero coordinates ch i= 1,2,..., the set of all x) = (cixi), where x = (xi)eX again forms a sequence space /i, which we call adiagonaltransformofl Its a-dual jux is obtained from Xx by making the diagonal transformation by the vector b =1 — 1. The mapping s-»t) is clearly a topological isomorphism of X[X] onto ^[S]. The £-equicontinuous sets {u}" in Xx correspond
3. Sums and products of sequence spaces 409 to the £-equicontinuous sets {o}", where 0 = 1 — 1. We now assert: (4) The normal topology X of a sequence space X => cp is always coarser than the Mackey topology %k(Xx). The topological dual of a sequence space X[X] => cp is therefore always the same as its a-dual Xx. It is sufficient to show that every set (u}"c=Ax is £s(A)-compact. Since such a set is certainly absolutely convex, we shall then have $1 <= ft, where ft is the collection of £fc(A)-equicontinuous subsets of Xx. By the remarks made above, we can restrict our attention to the case where u = e = (l,l,...), provided that all the coordinates of u are nonzero. But then X is a subspace of I1. Since {u}" is the unit ball of /°°, and is therefore ^(/^-compact, {u}" is a fortiori £s(A)-compact. If we delete coordinates on which u vanishes from the vectors of X and of Ax, we reduce the general case to the one which we have just considered. 3. Sums and products of sequence spaces. If the coordinates of the vectors of a sequence space X are permuted, so that x = (x1,x2,...) is transformed to x' = (xni,x„29...), we obtain a space fi[%] which is topologically isomorphic to /[£], and whose dual is obtained by making the same permutation of Xx. For since £w,-xf is absolutely 00 convergent for ug/x, xe/, we have U£ = u'£' = £ un.xnr i= 1 Sequence spaces which are transformed into themselves by any permutation are said to be symmetric. Examples are (p,a> and the /^-spaces, lg/?=oo. We can also rearrange the coordinates in another way, where, instead of using the positive integers, we use any countable partially ordered index set. The vectors can, for example, be written as double sequences. If we make the same rearrangement of the a-dual, we again obtain a sequence space which is topologically isomorphic to the original one. GO If XX,X2,... are sequence spaces, then the cartesian product TT Xh which consists of all double sequences (x(1),x(2),...), x(i)eXh is again a sequence space, in the sense which we have just described. In the same 00 way the direct sum © Xt consists of the double sequences (x{i\...,x{n\ i=l o,o,...), where o stands for the zero-vector in X{n+1\ A(" + 2),..., and n is an arbitrary positive integer. It is not difficult to see that
410 §30. Perfect spaces (1) If the spaces Xt=> cp are normal (respectively perfect), then the 00 00 spaces © X{ and TT Xt are also normal (perfect). / oo \x oo / oo \x oo For arbitrary X{ =) cp we have I © X{ \ = TT A? and I T\ XA = © A*. v=1 / i=1 v=1 / i=1 If, further, the spaces X{ are given their normal topologies %, we have 00 00 (2) The normal topologies of © AipX] and TT A£[3f] are respec- i=l i = 1 £iWy £/ie topologies of the locally convex direct sum, and of the topological product. This is easy to prove either directly or by using § 22, 5. (3) and (4). By starting from cp and co, and repeatedly forming sums and products, we obtain the spaces of countable degree described in § 13, 5. These are all perfect spaces, and their normal topology is the same as the strong topology, since this is the case for cp and co, by 2., and since this property is carried over to the spaces formed from them, by § 22, 5. If nl<n2<"m is a sequence of positive integers, and if for each vector x = (x1,x2,...) of a sequence space X we form the vector x' = (xni,x„2,...), we obtain another sequence space \i, which we call a sectional subspace of X. Clearly, (3) If X^.(p is normal or perfect, this is also the case for any sectional subspace. The dual of a sectional subspace is obtained by leaving out the same coordinates of Xx. \x and the sectional subspace p! defined on the set complementary to {n1,n2,...} together form a complementary decomposition of X: X = fi®jj,'. 4. Unions and intersections of sequence spaces. A linear subspace of a sequence space is again a sequence space, whereas the quotient by a closed linear subspace need not be a sequence space. Two important ways of constructing new sequence spaces from given ones are special cases of forming hulls and kernels. If {Xa} is a family of sequence spaces, all of which are considered as subspaces of the same space co, the linear span £xa in a> is again a a sequence space. Similarly the intersection f]Xa in at is a sequence space. a. (1) a) // all the spaces Xa contain cp and are normal, the same holds for Y,K and f]K- a a b) The intersection Q Xa of perfect spaces Xa is perfect.
4. Unions and intersections of sequence spaces 411 c) // all the spaces Xa contain cp, we have w^K) = 0^« > tf ^e ^ a a. spaces Xa are perfect, we also have \C\^<x) =\YjK) * *• a a Proof, a) and c) follows easily from the definitions, b) follows from c): if the spaces Xa are perfect, (£Aax)x = f)X* x = f]Xa, and so f]Xa is perfect, as it is an a-dual. The following example shows that the equation (PUJ^L^a* nee<^ not always hold, even for perfect spaces Xa (cf. § 13, 5.): Let Xl = (pa> and 22=coc/>, where both are considered as subspaces of the same space co, written as a space of double sequences coco. Then Xx nl2 is equal to <p<p, the space of double sequences with only finitely many non-zero terms. Thus (Xx r\X2)x =coco. On the other hand, since (<pco)x = co<p, and (axp)x = <pco, X\ +X$ = coc/> + <pco, and this is a proper subspace of coco, as can be seen directly. For the finer theory of sequence spaces there is a special representation in terms of intersections which is particularly important, and which we shall now develop. Let a = (at) be an arbitrary coordinate vector. We denote the perfect 00 space consisting of all those xeco which satisfy £ \at\ Ix^oo by Xa. r= 1 If all but finitely many a{ are non-zero, Xa is a diagonal transform of I1, as we have already seen in 2. If there are only finitely many non-zero ai9 Xa is equal to at. If both infinitely many at are non-zero and infinitely many at are zero, Xa divides into two sectional subspaces, of which one is a diagonal transform of ll, and the other is equal to co. It follows from 1. that the a-dual Aax of Xa in general divides into two sectional subspaces, of which one is a diagonal transform of /°°, and the other is equal to cp. Aax is the perfect cover of the sequence space consisting of the vector a and its scalar multiples. We observe that for positive at we clearly have the relation k (2) Xai+...+ak = 2^ Kj- Every perfect space X consists of all those xeco for which Yj\ui\ \xi\<co f°r all positive ueXx. From this, and from (2), we get (3) Suppose that X is perfect. Then X = f] Xu and Xx = [J Aux, where u u u runs through all the positive vectors in Xx. If k=> <p is normal, X = [jX*, where x runs through all the positive vectors in X. * Using the terminology of § 19, 6., we can also express the fact that X = f]Xu by writing X=KIiu~1)(Xu), where Ju is the embedding of X
412 § 30. Perfect spaces in Xu. If we equip each space Xu with its normal topology, then we can give X the kernel topology, as in §19,6. It follows directly from the definition of this topology and from the definition of the normal topology on X that these two topologies on X are the same; thus it follows from (3) that (4) Every perfect space X[X], where X is the normal topology, is the locally convex kernel KI{u~l)(Xu[X]) of the spaces XU[X] defined u by the positive vectors u in Xx. If u and o are positive vectors in Xx, we set u<o if o — u>o. In this way the positive vectors u in Xx form a directed set. If for u < o we define Iuo to be the embedding of Xv in Xu, then Iuv is a continuous mapping from XV[Z] into XU[Z]. Using § 19, 7.(6) and § 19, 8.(1), we obtain the following stronger form of (4): (5) Every perfect space X[X] is topologically isomorphic to the projective limit lim/U0(ADpX]), where u and v run through all pairs of positive vectors in Xx satisfying u<o. The most important properties of sequence spaces can now be obtained from the topological properties of the spaces /U[3T|, and from the general theory of locally convex kernels and projective limits. 5. Topological properties of sequence spaces. As before, let ef denote, the vector all of whose coordinates vanishes except the z-th, which is equal to one. The n-th section of a vector x = (xi)eX[(X] is the vector n *n= Z ^e. = (Xl,...,x„,0,0,...). ; = i 00 If follows directly from the convergence of ux= £ uixi f°r eacn ueXx that I = 1 (1) The sequence x„ of n-th sections of a vector x in k"=> cp converges weakly to x. The mapping which is obtained by sending each vector xeX to its z-th coordinate x{ is given by eix = xi; since eieAx, this mapping is weakly continuous, and a fortiori it is a continuous linear functional on X for the finer topologies <Zk(Xx) and £. (2) // a sequence x{n) of vectors in a sequence space X is weakly *£- or ^-convergent to x(0)eX, x{n) is coordinatewise convergent to x{0\ For if x(n)->x(0) with respect to one of these topologies, it follows that efx(w) = xj-,l)->e,.x(0) = xj-0), by the remarks made above. A corresponding result holds for filters 5 which are Cauchy with respect to one of these topologies. If it exists, the limit x{0) of 5 nas> as z-th coordinate, the limit of the filter e^g-
5. Topological properties of sequence spaces 413 In the case where X = oj, all these modes of convergence coincide with co- ordinatewise convergence; this is the convergence which corresponds to the normal topology on co. We now give the first topological characterisation of perfect spaces. (3) A sequence space X => cp is perfect if and only if it is weakly sequentially complete. Xx x is obtained from X by taking the coordinatewise limits of the weak Cauchy sequences in X. Proof, a) Suppose that X is perfect, at is weakly complete, and a fortiori it is weakly sequentially complete. By § 22,4.(2), ll is also weakly sequentially complete; so therefore is each au, u>o, ueXx, by 4. From 4.4 it follows that (4) A[2j = lim/U0(AU[2J). This can either be confirmed directly, or we can use §22,7.(6). It then follows from § 19,10.(2) that X is weakly sequentially complete. b) Suppose that X => cp. By (1), xeXx x is the £S(AX x x)-limit of its sections s„, Since Xx x x =/x, and since xneX, x is the £s(Ax)-limit of the xneX. It follows directly from (3) and the Banach-Mackey theorem (§20,11.(8)) that (5) In every perfect space X the weakly and strongly bounded sets are the same, and so therefore are the topologies %b(Xx) and cXb*{Xx). The next result makes it easier to determine the bounded subsets of a sequence space: (6) If X^> cp and if X is normal, the normal cover of every bounded subset M of X is bounded. Since every bounded subset of X is also bounded in Xx x, and since X contains Mn if it contains M, we may suppose that X is perfect. Now in I1 and in at the normal cover of a bounded set is bounded. This therefore holds in each Xu, and so by 4.(4) and § 19,6.(7) it also holds in X. We now give a second topological characterisation of perfect spaces: (7) A sequence space X => cp is perfect if and only if it is %-complete, where X is the normal topology. Xx x is the %-completion of X^> cp. Proof, a) If X is perfect, /[£] is the projective limit of the spaces /u[3f|, by 4.(5). Since these are complete, X [X] is also complete, by §19,10.(2). b) As in (3)b), the fact that Xx x is the completion of X[Z] follows from the following result, which strengthens (1):
414 § 30. Perfect spaces (8) The sequence x„ of n-th sections of a vector x of a sequence space X => cp is H-convergent to x. It is sufficient to prove this for perfect X. The assertion is true in co and in ll, and so it is true in all the spaces ku, ueXx. But it follows from pu(x — £„)->0 for all ueXx that xn->x with respect to the normal topology on X. Since %k(X*) is finer that 2 (cf. 2.(4)), it follows from (7) that (9) Every perfect space is complete under its Mackey topology. If we restrict ourselves to normal spaces X containing cp, (8) can also be improved: (10) If k^xp and if k is normal, the sections xn of an xek converge to x with respect to the Mackey topology. This time we use the representation X = [JXX, xek, of 4.(3). X The assertion is true for cp, and it is also true for /°°; this can easily be seen from the form that the weakly compact subsets of ll take (§22,4.(3)). As a result, the assertion holds in each X*. But since Xx^> X, so that £fc(Ax) is coarser than Xk(Xx), the fact that xn->x with respect to Zk{kx) implies that xn->x with respect to %k(X*). (11) Every sequence space 2[1] d <p is sequentially separable. If further, k[X] is normal, k is also sequentially separable under the Mackey topology. n We shall show that the set N of vectors £ pf cf, with p{ rational and n arbitrary, are sequentially dense in k for the topologies £ and £k. n Given xek, we can find, for each n, an element x(n)= £ p|w)cf in N for which £=1 (12) |Xi_p(»)|^M9 i=i,...,w. n Now if B is a 3> or £fc-equicontinuous subset of kx, (13) SUp|u(3t-3t("))I^SUp|u(3t-3t„)| + SUp|u(3t„-3t("))|. ueB The first summand of the right-hand side is less than or equal to — for n^n0, by (8), and by (10), respectively. But the second summand can also be made arbitrarily small: B is a bounded set, and by (6) the normal 00 cover of B is bounded. Consequently sup £ \ut\ \xt\ = K< oo. It therefore follows from (12) that ueB <= * suplu^-s^suptk-l |x£-pj">| g - £ 4 n 2 for sufficiently large n; the assertion now follows, using (13).
6. Compact subsets of a perfect space 415 6. Compact subsets of a perfect space. Suppose that X is perfect and that X is a locally convex topology on X; the only assumption we make about X is that is finer than %S{XX). We have (cf. Kothe [9]): (1) Suppose that M is a subset of the perfect space X. The following properties of M are equivalent: a) M is X'-compact; b) M is countably X-compact; c) M is sequentially X-compact; d) M is bounded, and every sequence x(n)eM which is coordinatewise convergent to a vector x{0) in cd is X-convergent to xi0\ and x{0) lies in M. Proof, b) follows trivially from a). c) follows from b): Suppose that M is countably ^'-compact, and suppose that x(n) is a sequence in M. Since M is bounded, it is coordinate- wise bounded, and so using a diagonal procedure we can choose a coordinatewise convergent subsequence of the x(n). Let this be denoted by x(n) again, and let x0 be its coordinatewise limit. By hypothesis x(n) has a ^'-adherent point n0 in M. By the remark preceding 5.(2), however, this can only be the coordinatewise limit x0. For the same reason, x0 is the only possible adherent point of any subsequence of x(n\ so that it must be the S'-limit of x(n). This is essentially the method of proving Smulian's theorem (cf. §24,1.(2)). d) follows from c): Suppose that M is T-sequentially compact. M is then bounded. A sequence x(n) which is coordinatewise convergent to x0 can only have x0 as T-limit. If x0 were not the T-limit of x{n\ a subsequence would have to be convergent to some t)0=M0, and this is not possible. a) follows from d): Suppose that M satisfies hypothesis d), and that g={jpa} is a filter on M. If to each x = (xl,x2,...)eM we make correspond the element (x1,...,xn)eKn, we obtain from g = {Fa} a filter %n={F*} on K". Let Gn be the bounded set of adherent points of g„ in K". Let X)(n)eX be chosen in such a way that (y("\..., y^) belongs to G„. We can pick out a subsequence of the X)(n) which is coordinatewise convergent; this we again denote by X)(n\ Let its coordinatewise limit be xi0)eco. In each set Fa there is now an element 3(n) with \z\n) — y^ \ ^ —, for i = 1,..., n. n The sequence 3(n) clearly converges coordinatewise to x{0\ By d), x{0) belongs to M, and it is the T-limit of the sequence 3(n) in Fa. Thus x{0) is a ^'-closure point of each Fa, so that x(0) is a T-adherent point of the filter g on M. The fact that a) and b) are equivalent contains Eberlein's theorem for perfect spaces; in the present case, therefore, this can be proved in a much simpler way.
416 § 30. Perfect spaces The next result is important for determining weakly compact sets: (2) The weakly closed normal cover Mn of a weakly compact subset M of a perfect space X is always weakly compact. This is true for X = co, since every bounded subset of cd is weakly relatively compact. It is also true for X = l1, as can be seen from the structure of the weakly compact subsets of Z1 (cf. §22,4.(3)). Thus (2) holds for all the spaces Xu, ueXx. Now if M is weakly compact in a, it is a fortiori weakly compact in each space Xu. If xin)eMn is coordinate- wise convergent to x0, x{n) converges weakly to x0 in each Au, by (1) d), so that x0 belongs to X = f)/„u, and x(n) also converges weakly in X to x0. Mn is therefore weakly compact, by (1) d). (3) A subset M of the perfect space A[3f] (where 2, is the normal topology) is weakly compact if and only if it is %-compact. Weakly convergent and %-convergent sequences are therefore always the same. Suppose that M is weakly compact in X[X~\ = f]Xu{X']. M is weakly u compact in each space Xn. By §22,4.(3) M is strongly sequentially compact in /u, and a fortiori it is ^-sequentially compact in Xu. If x{n)eM is coordinatewise convergent to x0, x(n) is ^-convergent to x0 in each space au, by (1) d), and by 4.(4) x(n) is therefore also ^-convergent to x0 in X. Consequently M is ^-compact, by (1). Applying 5.(7), § 21, 7. and § 21, 9.(7), it follows from (3) that (4) The Mackey topology <Xk(Xx) on the perfect space X is the polar topology £° of the normal topology £ on Xx. Zk(X) is the finest locally convex topology which coincides on the %-equicontinuous sets of Xx with the weak topology. The polar topology of Zk can also be characterised in a simple way: (5) The topology 3£, the topology of uniform convergence on the ^-compact subsets of the perfect space X, is the finest locally convex topology X' on X* which gives the same convergent sequences as the weak topology. By §21,9.(7), 3£ is the finest locally convex topology on Xx which coincides with the weak topology on the weakly compact subsets of Xx. It follows from this that every weakly convergent sequence is ^-convergent. Conversely if every weakly convergent sequence is ^'-convergent, the weakly compact and ^'-compact sets are the same, by (1), and so Z°k is finer than %'. By (3) and (5), ££ is always finer than Z. The next result is analogous to (2): (6) The weakly closed normal hull Mn of every %k-compact subset M of a perfect space is again Zk-compact.
7. Barrelled spaces and (M)-spaces 417 If Mn were not ^-compact, there would be a weakly convergent sequence u(n)-> o in Xx for which sup |u(n)x|>m>0, by (5). There xeiVT' oo would thus be a sequence x(n)eM with £|t4n)| |xjn)|^m. Since u(n) i converges weakly too, every sequence v(n) with |ujn)|= |w|n)| converges 00 weakly to o as well, by (3). If v(n) is chosen in such a way that XlwS"}l lx/n)l i = v{n)x{n\ sup |o(n)3e| would not converge to 0, in contradiction to (5). xeM The normal cover of a strongly compact subset of a perfect space need not be strongly compact, as the example of the set consisting of the vector c = (l, 1,...) in /°° shows. 7. Barrelled spaces and (M )-spaces. We saw in 5.(10) that in every perfect space X the sections xn of a vector x converge to x in the topology 3fe(/x). As the example of /°° shows, this need not be so for the strong topology. The class of perfect spaces for which the sections converge strongly can be characterised in the following way: (1) The following properties of a perfect space are equivalent: a) A[£fc] is barrelled; b) ^-convergent and strongly convergent sequences are the same; c) every ^-compact subset of a is strongly compact; d) the weakly closed normal cover of every strongly compact subset of a is strongly compact; e) the sections of every xea converge strongly to x; f) a [Ift] is sequentially separable. Proof, b) and c) follow from a), and are equivalent, by 6.(1), d) follows from c) by 6.(6). If d) holds, the weakly closed normal cover of the set consisting of a single vector xea is strongly compact. From this it follows by 6.(1) that the sections of x converge strongly to x, so that e) holds, f) follows from e) as in 5.(11). In order to be able to derive a) from f) we must show that every weakly closed bounded subset M of Xx is weakly sequentially compact. Let u(n) be a sequence in M, and let xil\ i= 1,2,..., run through a strongly sequentially dense sequence in a. By using a diagonal procedure, we obtain a subsequence, which we again denote by u(n\ for which limu(n)x(0 exists, for each i. We therefore have (2) Ku^-u^)^0!^! for n,w^«o(fi,0. Let xe/, and let xU) be a subsequence of x{i) which converges strongly to x. Since the set of all expressions u(n) — u(m) is bounded in ?.*, there is &j0 for which 27 Kothe, Topological Vector Spaces I
418 § 30. Perfect spaces (3) \(u{n)-u{m))(x-xij))\^^ for all n,m and for j^j0{e). From (2) and (3) it follows that |(u(n)-u(m))x|^e for n9m^n0(ej0). Thus the u(n) form a weak Cauchy sequence in M, whose limit lies in M, by hypothesis, and by 5.(3); thus M is weakly sequentially compact. Since a perfect space X[X~\ is semi-reflexive if and only if (Xx [3^(A)])' = A, and so if and only if Xx \%k{X)~] is barrelled, the next result follows directly from (1): (4) A perfect space X[X] f w/zere £ is the normal topology J is semi- reflexive if and only if Xx satisfies one of the conditions a) to f) of (1). Likewise, we have (5) A perfect space A[£fe] is reflexive if and only if both X and Xx satisfy one of the conditions a) to f) of (1). For the Mackey topology, we have (6) // X is a perfect space, the following properties are equivalent: a) weakly convergent and %k-convergent sequences in X are the same; b) every weakly compact subset of X is Hk-compact; c) the topologies Zk(X) and 3£(A) on Xx are the same. A perfect space X has these properties if and only if its a-dual Xx has them. Proof, a) and b) are equivalent, by 6.(1) a) and c). Further b) is equivalent to c). By 6.(5) the weak topology and the topology 3£ coincide on a weakly compact subset of X*. Thus if c) holds, every weakly compact subset of Xx is ^-compact, i. e. b) is satisfied, for Xx. Let us remark that in (6) weak convergence and the weak topology can also be replaced by normal convergence and the normal topology, using 6.(3). In particular X has the properties described in (6) if the normal topology X is the same as the Mackey topology. (7) Weakly and strongly convergent sequences in a perfect space X are the same if and only if the bounded sets of Xx are relatively %k-compact. Such a space is always barrelled. This follows directly from (1) and (6). We now give a criterion for a perfect space to be an (M)-space. We observe that, by (1), every perfect (M)-space is sequentially separable. (8) A perfect space A[£fe] is an (M)-space if and only if the sections of each vector x in X converge strongly to x and the weakly ( or %-) convergent sequences in Xx are the same as the strongly convergent ones. By (1), the first condition means that A[Ifc] is barrelled and that every ^-compact subset of / is strongly compact. The second condition means that every bounded closed subset is ^-compact, by (7). The two together give the result.
8. Echelon and co-echelon spaces 419 T. and Y. Komura [1] have given an example of a barrelled perfect space which is not bornological. 8. Echelon and co-echelon spaces. In 4. we introduced the special perfect spaces Xa and their a-duals A*. If we are given countably many vectors a(fe), fc=l,2,..., in co, which we call steps, we call the linear 00 00 span Y, Kw tne co-echelon space, and the intersection f] AaW fc=i fe=i the echelon space, corresponding to them. We know (cf. 4.(1)) that f] XqW is perfect, and that £xax(k) *s normal. The steps can all be supposed to be positive. By possibly going over from the steps a(fe) to the steps a(1) + *-- + a(fc), we can obtain a mono- tonic increasing system of steps, without altering the spaces. Further we can suppose that the system of steps is complete, in the sense that for each index i there is an a(k) for which a^ is non-zero. In what follows we shall always make these assumptions. Then £Aax(k) is the union of the spaces Aax(i) c Xx{2) <=•••. The projective limit topology X' on / = P)Aa(k) is defined by the oo k semi-norms qk(x) = £ |a-fe)| |xf|,fc=l,2,..., so that A[2'] is an (F)-space. ; = i The dual X of A[T] is equal to the co-echelon space \J A^k). X\ being k the dual of an (F)-space, is weakly sequentially complete. Since (X)x =x (cf. 4.(1)), X is perfect, by 5.(3), so that X= (J/.*(*>=/.*, and X' is the same as the normal topology X on X. Thus we have shown (1) The spaces f] Xa(k) and (J/laXk) are perfect and a-dual to each other. The normal topology X on X=f]Xa(k) is given by the semi-norms qk(x\and X[X~\ is an (F)-space. On any (F)-space, the topology is the strong topology. Thus it follows from X = Xb(Xx) and 6.(3) that weak and strong convergence of sequences coincide in every echelon space. Further, h[X] is sequentially separable (7.(1)). In addition it follows from X = Xb(Ax) that every bounded subset of the co-echelon space Xx =\J/.*ik) lies in the normal cover of a vector pa{k\ and so it is bounded in some Aax(/c). A weakly convergent sequence u(n)->u inAx is therefore bounded in some Xx{k), and is coordinatewise convergent. It is therefore also weakly convergent in some Ax(k). The definition of co-echelon space was generalised by Dieudonne and Gomes [1]. Corresponding to a positive vector a, with at>0 for 00 all i, we form the set Ap of all x with £ \at\ \xt\p< oo, for some p^i 1. i = 1 AJ, equipped with the norm ^(x) = (^|af| |xt|p)1/p, is a diagonal trans- 27*
420 § 30. Perfect spaces form of /p, and so it is perfect. Its dual (and at the same time its a-dual) is the space (Xp)x of all u with £|af| ^|mi.|«<oo, - + -= 1. {Ap)x is a diagonal transform of lq, and has norm q(u) = (YJ\ai\ p\ut\q)q. If infinitely many at are zero, Xp decomposes into a sectional subspace which is a diagonal transform of lp and one which is isomorphic to o; corresponding to this there is a sectional subspace cp of (Xp)x. Once again let a(1)^a(2)^- •• be a complete sequence of positive 00 steps. The intersection f] Xp{k), with the topology Z0 given by the fe=i semi-norms qk(x) = (YJ\aiik)\\xi\p)l/p, fc=l,2,..., is called the echelon space of p-th order corresponding to the a(fe); it is an (F)-space. 00 Similarly, the space (J (Xp{k))x is called the co-echelon space of n=l p-th order. When p=l, we obtain the echelon spaces considered above. Using the same method of proof as before, we obtain 00 00 (2) The spaces f] Ap(k) and [J (Ap{k))x are perfect and a-dual to each other. fe=1 fe=1 The topology Z0 = Zb(Xx) is of course in general strictly finer than the normal topology for p>l. However X = f)Xp(k) is always sequentially separable. It follows from the fact that Z0 = Zh that every bounded subset of the co-echelon space ax = (J(^(k>)x *s contained in a scalar multiple of the unit ball of some (AJ(kj)x, and that weak sequential convergence in ax means the same as weak sequential convergence in a suitable (Apik))x (for a bounded coordinatewise convergent sequence in /p, p > 1, is weakly convergent in lp). It follows from §23,3.(7) and the fact that the spaces Xp(k) are reflexive for p > 1 that (3) Every co-echelon space Ax[Ife(A)] and every echelon space ^[£0] of order p>\ is reflexive. Example. If the vectors a{k) = (\,k,k2,...), k=\,2,..., are taken as steps, the corresponding echelon space X can be considered as the space of entire transcendental functions: to the vector x = (x0,xl,...)el we assign the function 00 x(z) = Yj xiz*- If. m a corresponding way, we assign to each ue/x the function u(z) = -1 u0 + wx - + u2 ~2 +' • •), we obtain the space of all functions which are analytic in a neighbourhood of oo and which vanish at oo. The scalar product can now be interpreted in the following way: 1 r °° (4) -—:&>u(z)x(z)dz= Y UjX: = ui. 2nlJ i = o
9. Co-echelon spaces of type (M) 421 For on multiplying the two series together we obtain the Laurent expansion of u(z)x(z) in an annulus about 0, and ui is the coefficient of 1/z. The integral is taken along a circular contour about 0 which lies within the annulus. In the terminology of §27,4., H(T) is the space of entire functions, with the topology defined there. Using (4), we obtain an isomorphism of the dual pairs <H(oo),H(0> and </lx,/l>; in particular X[%~] and Xx [Xfc] are topologically isomorphic to H(V) and H(co) respectively, when these spaces are given the topologies defined in §27,4. Consequently X[X~\ and /lxpXJ are (M)-spaces, each being the other's dual. For this, cf. Toeplitz [2]. 9. Co-echelon spaces of type (M). A precise criterion can be given for a co-echelon space to be an (M)-space. We have (cf. Kothe [6], Dieudonne and Gomes [1]): (1) Suppose that a(fc) = (ajk)), fc=l,2,..., is a complete monotonic increasing system of steps, and that A[I0] and ^x[£fc] are the corresponding echelon and co-echelon spaces of order p^l. These two spaces are both (M)-spaces if and only if there is no infinite index set {jn} such that for some suitable fc0, and for suitable Mfe>0, we have (2) 0<a£>^Mka£o) for all k^.k0. Expressed in another way: X and Xx are both (M)-spaces if and only if there is no sectional subspace of X ( respectively X x) which is a diagonal transform of lp ( respectively lq) I - + - = 1, q = oo for p = 1 I. \P Q J An immediate corollary is: An echelon or co-echelon space of order 1 is reflexive if and only if it is an (M)-space. Proof of (1). By 7.(8), / is an (M)-space (and so, therefore, is Xx) if and only if weak and strong sequential convergence coincide in Xx. This does not happen if Xx has a sectional subspace which is a diagonal transform of lq, q> 1, or of /°°. We must therefore show that we can deduce (2) from the fact that there is a sequence u{n) in Xx which converges weakly but not strongly to o. By 8., we can suppose that u(n) converges weakly to o in some (AJ(ko))x. By making a diagonal transformation, and restricting attention to the sectional subspace defined by the non-zero a\ko\ we may suppose that a(feo) = e = (l,l,...), so that (AJ(ko))x =lq. Suppose therefore that \\u{n% = (YJ\u(in)\q)1/q=l, and that u(n) converges coordinatewise to o. Further suppose that M is a normal bounded subset in X for which (3) supf;Mn)lk-|^2c>0 for all n. "M' = 1 For each u(n) there therefore exists an x(n)eM for which C<ZM")lW",|^l|U<",IU|X(-)||p=||3E(")||p.
422 § 30. Perfect spaces From this, from the fact that M is normal and from the fact u{n) converges coordinatewise to o it follows that there exist infinitely many vectors ^"^(O,...^,^,...,^^^,...) in M, with \\x)(n)\\p>c and sn<rn + l for n=l,2,.... Since M is bounded, there exists mk<co such that 00 (4)' (&(*))"= EkflW^m*, k=\,2,..., i = l for all xeM. 00 | J Suppose that dk>0 and ^- = -. Let £fcll run through all indices rn^tkn^sn for which k = °dk 2 mudu (5) a[k?>^~±. For the corresponding coordinates of n(n) (with k,n fixed) we have (6) Ib£ilp<?> mti c p since otherwise we would have Xla!kil l>7SkilP>~V^'T" = mfe' which would contradict the fact that n(n)eM, and (4). If we form the sum over fc0,fc0+l,..., it follows from (6) that 00 1 cp 1 Zb£lp<cPZ- = -, for each n. k=kQtkn ak i But ||t)(n)||^= £ \yf)\p>cp. Consequently for each n there is at least one index jn with rn^jn^sn which is different from all the indices tkn, k = k0, fc0+l,.... Then for this7M (5) is false for all fc^fc0, i.e. we have afJ^"^ = Mk. Thus (2) is proved, with a(ko) = e. In §28,3. we defined local convergence. In every (F)-space, and in particular in every echelon space, local convergence is the same as topological convergence. We now determine those co-echelon spaces (of order 1) in which weak convergence is the same as local convergence. (7) Suppose that a{k) = (a{k)), fc=l,2,..., is a complete monotonic increasing system of steps, and that Xx is the corresponding co-echelon space of order 1. Weak convergence and local convergence coincide in X x if and only if for each k there exists an N(k) such that a{k) (8) !™^=0; here i runs through those indices for which a{fe)4=0.
10. Further investigations into sequence spaces 423 Proof, a) Suppose that (8) is satisfied and let u(n) be a sequence in Xx which converges weakly to o. There then exists a k and an M such that |wjn)|^Ma[k), and limw[n) = 0, for all i and n. But then it H-+00 follows from (8) that supla^l-1 \u\n)\-+0 as rc^oo, i. e. that u(n) con- i verges strongly to o in X*iN(k)). But this means that u(n) converges locally to o. b) In order to show that (8) is necessary we need only assume that the sections u„ of each u in lx converge locally to u. The sequence untn must then converge locally to o. If in particular we take a(fe) as u, d^tn must converge strongly to o in some suitable /.*<N<k». (8) follows from this. By (1), every co-echelon space which satisfies (8) is an (M)-space; conversely the system of steps of an (M)-space always satisfies (1), but need not always satisfy (8). If Xx is a co-echelon space which is an (M)-space but which does not satisfy (8), there must always be vectors u in Xx whose sections converge strongly to u, but which do not converge strongly to u in any X*(k). In § 31, 5., we shall discuss such an example in more detail. Every co-echelon space Xx [£fe] of order 1 which is an (M)-space is bornological, by §29,4.(4), and so it is the topological inductive limit of the spaces ^0Xk)[3^], where Xb is the norm topology. It follows directly from (8) that every bounded subset of X * [£fe] is relatively compact in some suitable a0xn)[3^]. The compact subsets of Ax[Ifc] therefore coincide with the sets which are compact in some A0xn)[IJ. If (1) is satisfied, but not (8), X* [Xk~] contains closed and bounded, and therefore compact, sets which are not compact in any X*(N)[Zb~\. 10. Further investigations into sequence spaces. We have considered sequence spaces with their structure determined by the a-dual Xx. It is also possible to follow rather different lines. For example, one can introduce the /?-dual Xp of a sequence space /, which consists of all those veco for which the scalar product 00 oi= Y, vixi converges (not necessarily absolutely) for all xeX. Xfi^X*, and we i=l now obtain the dual pair </r,/l>. This extension leads to rather complicated questions, however, as was pointed out by Kothe and Toeplitz [2]. In recent times this idea has been taken up again by Chillingworth [1], Matthews [1] and Garling [2], [3]. The theory of perfect spaces has only been developed here to a level which enables us to produce some examples which are important for the general theory. For further results reference may be made to the work of Toeplitz and the author, mentioned above, to the works of Allen, Cooke and their school (for this, cf. Cooke [1], [2]), and to the further works of the author. In two works which have recently appeared (T and Y. Komura [1], Pietsch [1]), further important developments have been made to the theory, and Pietsch [1] has generalized it to spaces of sequences whose terms are elements of an arbitrary locally convex space.
424 § 31. Counterexamples Another general class of sequence spaces was introduced by Zeller [1]. A sequence space I is called an (FK)-space if a locally convex topology X is defined on a, under which I is an (F)-space; further it is required that each mapping a = (ak)-*ak is a continuous linear functional on ApX']. Thus if a sequence a{n) converges to a, it must also converge coordinatewise. Above all, this class of sequence spaces has found applications to problems in function theory, and to the theory of summability (cf. Zeller [1], [2]). The theory of perfect spaces has been carried over to spaces of functions by Cooper [1] and Dieudonne [7]. Details of this will be given in the second volume. § 31. Counterexamples 1. The dual of/00. Our aim is to show that c0 is a closed subspace of /°° which has no topological complement. We begin by giving a representation of the dual of /^°, where d is any infinite cardinal. Let / be the set of indices on which the elements x = (£a), ae/, of /^ are defined. We shall write /°°(7) instead of /". If M is a subset of /, let /°°(M) be the sectional subspace of /°°(7) defined by M. If u is a continuous linear functional on /°°(/), let uM denote both the restriction of u to /°°(M), and the continuous linear functional on /°°(7) which is obtained from it by assigning the value 0 to the elements of /°°(/~M). (1) // Mu...,Mn are pairwise disjoint subsets of I, (2) IKJI^IMI- i=l For if elements xt are chosen in /°°(Mf) in such a way that ||xf||^ 1 „ n n and w^^g: ||wM.||—, £= Yj xi nas norm 11*11 = 1> and ux= £ uMixt n ;=i i=i ^ZIIwm,II — e> fr°m which (2) follows. Let 3 be the collection of all subsets M of /. Let eM be the element of /°°(7) whose coordinates are 1 on M and 0 on the complement of M. eM is thus the characteristic function of M. For each we(/°°)', we now define a set function <p(M) on 3 with values in K, by setting (3) <p(M) = weM, MeZs. If we define (a^+a2(/>2)(M) to be olx (p1(M) + (x2(p2(M), the mapping u-^cp is clearly linear. The set functions cp{M) are finitely additive, i.e. we have (4) (p ( U Mt J = Yj <P(Mi), for pairwise disjoint Mf.
1. The dual of/00 425 This follows directly from (3). If we define the variation of cp to be (5) K(<p) = supf>(M;)|, ; = i where the supremum is taken over all systems of finitely many pairwise disjoint subsets Mx of/, it follows from (2) that K(<p)^||w||<oo. We denote by B V(I) the space of all finitely additive set functions (or measures) cp{M) on /, with the norm given by (5). (6) The dual of /°°(7) is norm isomorphic to BV(I). n Proof. The collection of linear combinations £ aIeM.? with pair- i=l wise disjoint Mh forms a dense linear subspace H of /°°. From this it follows that the mapping u-*cp is a one-one mapping of (/°°)' onto a linear subspace of BV(I). Conversely, if cp is an element of BV(I), the equation (7) ul X a;eMi.j= £ ai<p(Ml defines a linear functional on H. If £a;eMi. has norm less than or equal to 1, |af|^l for all f, and so "( E a«eMi) ^ E WiWviM^Viv). Thus each cpeBV(I) defines a continuous linear functional on H, and so it defines an element we(/°°)', with ||w||g K(<p). Since conversely we have K(<p)^||m||, (6) is proved. The set-function cp corresponding to u is also written as (8) ux = §xdcp; i the right-hand side is defined for xeH by (7), and is defined for arbitrary ace/00 by taking limits s(n)-»s, where x(n)eH. A corresponding result also holds for £° (cf. Hildebrandt [1], Fichtenholz and Kantorovitch [1],Yosida and Hewitt [1]). Let c0(I) be the space of all null-sequences in /°°(7). Then c0(I)' = l1(I). Thus if we(/°°)', the restriction u of u to c0(I) is given by ux = Yucol£>v where ^|ca|<oo. But this expression also defines a con- a tinuous linear functional on the whole of /°°, which we again denote by u. In this way we obtain the canonical norm-isomorphic embedding
426 §31. Counterexamples of /*(/) in l1{I)"=lco(I)'. We set u = ii + fi, and in this way obtain an algebraically complementary decomposition (9) (/»)' = I1 (/)0 co(l)1 of (J00)'. In order to show that this is a topological decomposition, we prove the following result of Dixmier's [1]: (10) Suppose that E is a (B)-space. There is a continuous projection P of norm 1 from E" onto F. Proof. If to each ueE" we assign its restriction u = Pu to E c F, P is a projection of E" onto E. Since the norm of u in E is less than or equal to the norm of u in F", P has norm 1, and is therefore continuous. (11) The decomposition (9) is continuous with respect to the norm topology of (J00)'. In (10) we put E = c0(I). The projection P of (J00)' onto I1 is then continuous, and has null space Cq. The assertion now follows from §15,8.(1). Every ux has the form ws = £ca£a, with £|ca|<oo; the corresponding measure is given by q>(M)= £ cp. If M, are countably many (ieM (00 \ 00 £=1 / i=\ i.e. the measure cp is cr-additive. Conversely if / is countable and if a cpeB V(I) is cr-additive, it follows easily on applying (2) to the one-point sets that the corresponding linearly functional lies in I1 (I). Every linear functional uec^ vanishes on c0, and so the corresponding measure cp vanishes on all the finite subsets of /, and is not cr-additive. 2. Subspaces of /°° and Z1 with no topological complements. We continue with the investigation of (/°°)'. (1) Let M1,M2,... be countably many pairmse disjoint finite subsets of I. For each we(/0O)/ and each e>0, there exists a subsequence M„k 00 such that the restriction uM of u to /°°(M), where M= [j M„k, satisfies the inequality \\uM\\^s. k=1 00 For \J Mt can be decomposed into countably many pairwise i=l disjoint sets M(l) of the form M, and it follows from 1.(2) that 00 Yj IIwm(«) II = llwll > tne assertion follows from this. i= 1
2. Subspaces of/00 and Z1 with no topological complements 427 (2) If u{n) is a sequence in /°°(iy, and if the Mt are defined as in (1), there exists a subsequence Mjk of the M- for whose union N we have u^ = u{S\foralln. Proof. By (1), there is a subsequence of the M£ on whose union M(1) we have Hi^cnll^ 1. Applying (1) again, there is a proper subset M(2) of M(1), consisting of certain Mh on which ||w$2>||^l and ||w(^|2)||^j. In the general case, there is a proper subset M{n) of M("_1) for which ||m($(„)||^— , for k^n. Let N be a set which contains exactly one M. n from each M{n)~M(n + 1\ We form /°°(JV). If x is an arbitrary element of /°°(JV) with ||s||^l, and ifx(n) is the section ofx defined by Mjtu* -uMJn, then luffix — x{n))\ ^— for n^k. Thus for each ace/00, u^\x) is given by oo n £ v{£ £ar, v{£ = u{$ ear, where the ar are the countably many indices in N. r=l But then the vector (i/afc)) which represents uff lies in ^(N). (3) // the sequence w(")e(/0O)/ converges weakly to o, the sequence uin) of restrictions to c0 converges to o with respect to the I1-norm. Proof. Let us suppose that this is not the case. By choosing a subsequence of the u(n\ if necessary, we obtain a sequence M1,M2,... of pairwise disjoint finite subsets of / for which (4) I l«(n)ej^8, aeMn for all rc, and for some £>0. By (2) there is a subsequence Mjn on whose union N we have u^ = u^\ for all n. By hypothesis the u^ell(N) converge weakly to o, but by (4) they do not converge strongly to o. But this contradicts the fact that the weakly and strongly convergent sequences in I1 are the same. (5) There is no topological complement of c0(I) in /°°(/). Let us suppose that there is a norm complementary decomposition (6) r = c0®H. Let en be the sequence of unit vectors in Z1 = c'0. We extend each e„ to an element en of (Z00)', by putting en equal to 0 on H. Because of (6), en is £s(/°°)-convergent too in (Z00)'. The restrictions en = e„ to c0 are however not strongly convergent to o in J1, contradicting (3). (3) and (5) are due to Phillips [1]. For the present account, cf. Bourbaki [6], II, p. 118. Fichtenholz and Kantorovitch [1] proved the analogous result that C[0,1] has no topological complement in L°.
428 §31. Counterexamples It follows from (5) that the ^(/°°)-complementary decomposition 1.(9) is not £s(/°°)-complementary. For if this were so, then by § 20, 5.(1) there would be a corresponding ^((/^-complementary decomposition ;°° = c0 © H. c0 and H would then be closed subspaces of /°°, and we would have a contradiction to (5), by § 15,12.(6). It also follows from this that the projection P of (I00)' onto I1 is £fo(/°°)-continuous, but is not £s(/°°)-continuous. It also follows immediately from 1.(10) and (5) that c0 cannot be the dual of a (B)-space; this we proved in a different way in § 25, 2.(7). It is very easy to give examples of closed linear subspaces of I1 which have no topological complement. Let £ be a separable (B)-space in which there are weakly convergent sequences which are not strongly convergent. The spaces lp, p>l, provide examples. By §22,4.(1), E is topologically isomorphic to a quotient space ll/H. If H had a topological complement in Z1, this would be topologically isomorphic to E. But then weak and strong convergence would have to be the same in £, since this is the case in J1. 3. The problem of complements in lp and LP. The question of whether there are closed linear subspaces without topological complements in the spaces lp and LP, p> 1, p=l=2, was originally raised by Banach, and a positive answer was first given by Murray [1]. We shall give a rather simpler construction, due to Sobczyk [1]. A continuous endomorphism U of a locally convex space E[X] is called an involution if U2 = I, the identity. The set H of all x with Ux = x is called the subspace of the involution. H is always linear and closed. Projections and involutions are closely related. (1) If P is a continuous projection of E\jt] onto H, U = 2P — I is an involution with subspace H. Conversely if U is an involution with subspace H, P = j(U + I) is a continuous projection of E onto H. Proof. Clearly (2P-I)2 = I and \}{U+ l)~\2 = k{U+ 1). If Px = x, (2P — I)x = x, and conversely if Ux = x,j(U + I)x = x. (2) If P = j(I+U) is a continuous projection of E[%] onto H, then all the projections onto H are given by the mappings of the form P = j(I+U + V\ where V is a continuous endomorphism which satisfies (3) UV=-VU=V If P is a projection onto H, we must have PP = P and PP = P. The first condition implies that i(/+[/)(/+C/+K) = i(/+C/) + i(K+[/K) = i(/ + C/+K), so that UV=V.
3. The problem of complements in lp and LP 429 Similarly the second condition implies that —VU=V. Conversely it can be confirmed directly that every P whose V fulfils the conditions of (3) satisfies the equations PP = P and PP = P, and so, since PP = PpP = PP = P, it represents a projection onto H. If H is a closed linear subspace of the (B)-space £, let p{H) denote the infimum of the norms ||P|| of the projections P of E onto H. Similarly let u(H) be the infimum of the norms \\U\\ of the involutions U with subspace H. If there is no projection or involution on H, p(H) and u(H) are set equal to oo. (1) implies the following inequalities (4) \{u{H)-\)^p{H)^{u{H)+\). We now investigate the norms of projections in the rc-dimensional spaces lp. (5) Suppose that l^p=co, p=l=2, n = 2v. There is a linear subspace I1-1! f \--A\ \ Hin) of lp with u{Hin))= n\p 2\and p(Hin))^[n\p 21 - \). Proof. Because of (4), it is sufficient to prove the first inequality. We may restrict our attention to the case where p < 2. For if P is the projection onto H{n) and if U is the corresponding involution, then U' corresponds to P', and P' is a projection onto H(n)1. Since ||l/'|| = \\U\\, we have u(Hin)1) = np 2 = n\q 2| for H<n)1czlqn - + -=1 \P Q Let 2l1=r _|\ and generally let «v = Qv_1 _^v_1 )• Then the nxn matrix U = 2lv is orthogonal and symmetric, and so ]/n it defines an involution in lp. Let H{n) be the subspace of this involution. By (2), every involution with Hin) as subspace is given by a matrix U + 33, which satisfies the conditions of (3). If follows from (3) that the trace of 33 is equal to zero. Thus if VL = (uik) and 33 = (i;l7c), there must be at least one non- n negative vkk. Since U'U=(£, we have I = £ u2k. From 33 =11 33 and the symmetry of U it follows that l=1 l = \+vkk= J]uik{uik + vik] Since (i\uik\«J = 1 1 _ A q =n2 p, Holder's inequality gives A _ A l^H2^||(U+93)eJ|.
430 §31. Counterexamples But from this it follows that ||U+S||^n* 2, i.e. u{H{n))^np 2. The case p= 1 can be dealt with in the same way without any difficulty. Sobczyk showed that in fact if n = 2\ u(Hin)) = n\p q\ and that this value is the largest possible value for u(H\ for H <= lp. 00 If we form the space F= © Zfv, FcP. In the same way we can v= 1 00 construct the space H = © H{n) from the subspaces H{n) a lp2v. Then _ v= 1 H cz F, and H is a closed linear subspace of lp. (6) 7/ lrgprg oo, p=|=2, H /zas no topological complement in lp. By (5) it is sufficient to show that a projection P of lp onto H must, for each n = 2v, have a greater norm than any projection of lp2, onto H{n). Let g„ be the projection of lp onto J§v which sends the other spaces /fv', v' 4= v, to zero. Since Qn maps H onto H^, <2„P is a projection of lp onto H{n\ Its restriction (6„P)B to lp is a projection of lp onto /f(n). Since IIQJI^ 1 and since ||(6BP)J g |I2„P|| ^ ||P||, the assertion now follows. If p= oo, F is contained in c0, and the proof of (6) also holds for c0; thus there are also closed linear subspaces of c0 which have no topological complement. The next simple observation enables us to find further counterexamples : (7) If the locally convex space E has a closed linear subspace H with no topological complement, and if E is topologically isomorphic to a closed subspace of the locally convex space F, H does not have a topological complement in F, either. For the restriction to E of a continuous projection of F onto H would be a continuous projection of E onto H. From (6), (7) and §21,3.(6) it follows that C[/], /=[0,1], has a closed linear subspace with no topological complement. Following Banach [3], we have (8) // 1 Spt^ °o, lp is norm-isomorphic to a closed linear subspace of L'[0,1]. fl j Proof, a) p<oo. Let us put yi(t) = 2llp in —, -r—^ I, and zero /i y/p L2' 21 J elsewhere. Then ||^-|| = ( j Ij^dM =1, so that j^el/. The mapping \o / A which sends each x = (£>i)elp to the function x(t)= £ ^y^eU i=1 i is a norm isomorphism of V onto a subspace of LP, since \\x(t)\pdt oo 0 = £|&|", so that Mx|| = ||x||. i= 1
4. Complements in (F)-spaces 431 b) p=co. We follow the corresponding argument through, putting Vi(t) = l in —, ^—r , and zero elsewhere. \_2l 2I~1J It follows from (6), (7) and (8) that (9) Every space Z/[0,1], 1 ^ p ^ oo, p #= 2, has a closed linear subspace with no topological complement. Kakutani [4] has shown that if a (B)-space of at least three dimensions has continuous projections of norm 1 onto every closed linear subspace, then if must be norm isomorphic to a space Ij (cf. Bourbaki [6], II, pp. 142/4, as well). No infinite-dimensional (B)-space is known, other than the spaces I], and spaces topologically isomorphic to them, which has continuous projections onto all its closed linear subspaces. For further examples cf. James [3] and Komatuzaki [1], [2]. 4. Complements in (F)-spaces. Not only does every closed linear subspace of Ij have a topological complement, but so also does every closed linear subspace of the spaces cpd, cod, cpdl ©cod2, cpco and cocp, by § 12,1.(5) and Hagemann [1]. Let us remark, without proof, that this is also true for cp@l2 and co®l2; in addition Ornstein [1] has recently produced an interesting class of spaces with this property. At present these seem to be the only known examples of complete locally convex spaces with this property. We now obtain another negative result, this time for (F)-spaces. Suppose that E[%~] is an (F)-space. If E has an absolutely convex neighbourhood U of o which contains no straight line, the corresponding Minkowski functional is a continuous norm on E. Conversely, to every continuous norm on E there corresponds a neighbourhood of o with the given property. If £[£] possesses no continuous norm, X can be defined by a sequence of increasing semi-norms, no one of which is a norm. An (F)-space E[X] possesses a continuous norm if and only if there is a bounded 2s(£)-total subset of E'. co is an example of an (F)-space with no continuous norm; every space H((5), © a domain, is an example of an (F)-space with a continuous norm (cf. § 27, 3.). (1) Suppose that E[X~] is an (F)-space which is not a (B)-space, but which possesses a continuous norm. E has a closed linear subspace H with E/H = a>, and H has no topological complement. Proof. Using the terminology of § 29,1.(8), we consider a sequence vn e E'Bn ~ E'Bn_ 1. The vn are linearly independent; let G be their linear span. Since the sets Bn form a fundamental sequence of bounded subsets of £', only finitely many vn lie in each bounded set; the bounded subsets of G are thus finite-dimensional. By §21,10.(5), G is therefore a weakly
432 §31. Counterexamples closed subspaceofF. We set G1 = H, and form the space E/H. Z = Zk{E') and so by § 22, 2. (3) the induced quotient topology X on E/H is equal to <Xk(H1) = (Xk(G). Now the weakly compact subsets of G are finite-dimensional, so that X = Zk(G) = Xs(G) = Xs(H'). As a result the space E/H, which is complete under % is weakly complete, and so it is topologically isomorphic to cd. If B had a topological complement H0, this also would have to be topologically isomorphic to cd. But there would then have to be a continuous norm on H0, and this is not the case. We observe that we have established that every (F)-space which is not a (B)-space has a quotient E/H^co. Suppose that E is an (F)-space with a linear subspace H which has no topological complement, and which is such that E/H = co. Using the dual pair <£',£>, we can also introduce the linear weak topology Xls(E') on E. We now assert that E[%J provides an example of the circumstances described at the end of § 10,12. As was shown in § 20, 3.(2), the lattices of 2s(£')-closed and 2s(F)-closed linear subspaces of E are the same. Since H does not have a ^-complement in £, neither can it have a X/s-complement. The ^-isomorphism E/H = a> is also a ^.-isomorphism, so that E/H is linearly weakly compact. These observations are due to Dieudonne [10]. (1) is a generalisation of the following example of Toeplitz and the author (cf. Kothe [12]). Let a(z) be a holomorphic function with infinitely many simple zeros zk in the simply connected bounded domain (5. The operation A of multiplying the functions x(z)eH((5) by a(z) is a continuous endomorphism of H((5). Let F be the image-space. Since the uniform convergence of a(z)x„(z) implies the uniform convergence of x„(z), F is sequentially closed, and so it is closed; as a result A is a topological isomorphism of H(©) onto F. Let us determine FL a H(Q~ (5). We are therefore looking for all those functions w(z), holomorphic in some neighbourhood of Q~05, and vanishing at oo, which satisfy -—&>u(t)a(t)x(t)dt = 0 2mJ c for all x(z)gH((5). We may restrict our attention to the subset of all x(z) = 7, z —/ /gQ~(5, since this is total in H((5). We therefore require that (2) -L(f^df = o for all ,IeQ~<5; v 2m J t-X c u(t)a(t) is holomorphic in some annular domain (Q~(51)n©2, (51c=©2c= (5, which contains the closed curve C, and so it is the sum of two functions cy(t), which is holomorphic in (52> and c2(0, which is holomorphic in Q^©! and which vanishes at oo. Thus we have (3) liEMdt+l£cMdt = 0t ;.e^(5. 2niJt-X 2niJt-A c c The first integrand is holomorphic in (52, and so the integral vanishes. Considering the second term, it follows that c2(z) = 0 in Q~(5, and so also in
5. An (FM)-space 433 Q~©j. Since a(z)u(z) = cl(z) in (Q~ (51)n@2> we can extend w(z), which is holomorphic in Q—©^ analytically to all points of C52 which are not zeros of c (z) <z(z), by putting w(z) = ~Vt- Since <z(z) has only finitely many zeros zl5...,z„ a(z) in ©2, u(z) has the form £ . It follows from this that F1 consists of all k=1 z — zk finite linear combinations of the functions uk = , and so it has the form given in the proof of (1). z_Zfc It follows from (1) that //((5)/F = co, and that F has no topological complement. 5. An (FM)-space. We now resume the discussion of § 30,9. We start with vectors written as double sequences: a« = (flft>,ai*i,...;fl?1>,fl?l,...;...) = (b}k);...,bWi;kte;kt + 1e;...), fe=l,2 where bjk) = (l,2*,3*,...), c = (l,l,...). Using these a(k), we form the corresponding echelon space / and its dual co-echelon space /x. Since a(1) is the vector with <z-))=l for all i and;, Xx contains the space /°° = /aX(i), written as a space of double sequences. By §30,9.(1), A [I] is an (FM)-space, and Axplk] is its strong dual (M)-space. Xk(l) = Xb(?J on ax. It can be seen directly that the steps ak (once their coordinates have been arranged in a sequence) do not satisfy the conditions of § 30, 9.(8). Thus strong and local convergence in Ax are not the same. It is easy to see that a(1) is not the strong limit of its sections in any of the spaces /axk), whereas a(1) is the strong limit of its sections in Ax, by §30,7.(8). Thus the unit ball of Aaxi) = /00 is compact in >lx, but is not compact in any of the spaces A^k). Let A be the linear mapping from / which sends each x = (xij)ek (00 00 \ Z xfi» Z xi2>---)- Since a(1)e/x, it follows di- Z xu = ZZlxo"l<00' so that Axel1. rectly that Z j=U The adjoint mapping A' maps each ue/00 to the element (u;u; ...)gax ; this can be seen from the equations n(Ax) = Y,uJYdxij = (n;n;.^)x = {A'n)x. J i The linear subspace of all (u;u;...) is weakly closed in Xx; this follows from the form of the double sequences a{k\ A' is thus a weakly continuous one-one linear mapping from /°° into Ax with a weakly 28 Kothe, Topological Vector Spaces I
434 §31. Counterexamples closed image space. Using a theorem about mappings between (F)-spaces which will be proved in the second volume (cf. Bourbaki [6], Vol. 2, p. 106, Ex. 5 a), it follows that A is a topological homomorphism of /[£] onto J1. Thus a/N\_A] is topologically isomorphic to I1 in the topology induced by Xb(Ax). The (FM)-space a[£] therefore has a quotient space which is not an (M)-space. At the same time this example shows that a quotient of a reflexive (F)-space need not be reflexive. Let us now consider the canonical mapping K of a onto ?./N\_A]. Every closed bounded subset B of a is compact, and so its image K(B) is compact, too. But there exist bounded sets which are not compact in ?JN\_A]^ll; not every bounded subset of a/N\_A] is contained in the closure of the image K(B) of a bounded set. This means that the strong topology Zb(?./N\_A]) on H = N\_A]1 a ax is strictly finer than H is an example of a closed linear subspace of a barrelled space which is not barrelled, and is also not quasi-barrelled. Since a x \_<Xh] is bornological, H is an example of a closed subspace of a bornological space which is not bornological, by §28, 1.(1). Finally /x [%b~] is a (DF)-space whose closed linear subspace H is not a (DF)-space. This follows for example from § 29, 3.(12), since H is separable, but not quasi-barrelled. For this example, cf. Kothe [6] and Grothendieck [10]. 6. An (LB)-space which is not complete. Our example is constructed in a similar way to the example of 5. Let EQ be equal to c0, with the elements written as double sequences, * = (*11>*12 »••• J X2\^X22^'":> '•')' Thus x belongs to E0 if and only if lim |xl7c| = 0. E0 is a (B)-space under the norm ||3c||0=sup|xI-fc|. i,k Let d$ = k for i^n and for all /c, and let a{$=\ for i>n and for |x- I all k. Let En be the space of all double sequences with lim —^- = 0, | ''^ a* with norm ||x||n= sup—~- En is obtained from E0 by making a diag- i,k a\p onal transformation, and so it is topologically isomorphic to c0. The embedding of En_l in En is continuous. The topological inductive limit E = lim£„ therefore exists, provided that the hull topology H on E is Hausdorff, and E is then an (LB)-space.
7. An (F)-space which is not distinguished 435 be- I If we put aik = k for all i and /c, the norm ||*||00 = sup—— is weaker than ||i||„ on £„, and so it defines a Hausdorff topology on E which is weaker than X. Thus X is Hausdorff as well. It is easy to determine the dual space E. The dual E'n of En is a diagonal transform of Z1, and it consists of all those double sequences u 00 with Y, \uik\aiik <QO- Since the set of unit vectors cfk is total in each i,k = i En, and is therefore total in E, E' can be identified with the space of all 00 double sequences u for which £ lu^a^<co for w=l,2,.... i,fc = l Once again let e denote the vector (1,1,...). Let B be the set of all n vectors e(w) = (e,...,e,o,o,...), w=l,2,.... J5 is contained in E, and is GO weakly bounded, since |e(w)u|^ £ |wl7c|<oo, for each ue£'. Since e(M) i,k = l belongs to En, but not to En_u B is not contained in any one of the spaces En. If £ were quasi-complete, the closed absolutely convex cover of B, which would be complete, would have to lie in some £„, by § 19, 5.(5); a fortiori the same would be true of B. We have therefore shown that the (LB)-space E is not quasi- complete. Applying the final remark of § 19,5., it follows that we have also shown that the locally convex direct sum of countable many spaces c0, which is a complete locally convex space, has a quotient which is not complete. 7. An (F)-space which is not distinguished. We now take the sequence of vectors b^ = (^) = (CTib;e;e;...), fc=l,2,..., where b = (l,2,3,...) and e = (l,l,l,...), as the system of steps. By §30,8.(1), the corresponding echelon space a is a separable (F)-space under the normal topology. Its strong dual is the co-echelon space /x with the strong topology Xb(X). In order to show that /l[£] is not distinguished it is sufficient, by §29,4.(3), to show that /x [£fe] is not bornological. CO Let X be the hull topology of Ax = [J /bx(M), when each Abx„) is n= 1 given its norm topology. If Bn denotes the set of all uelx with GO \ui}\^b{V the sets V= \~ cnBn,cn>0, form a base of 2-neighbour- «= i hoods of o in Xx. We must show that X is strictly finer than Xb(l) on kx. 28*
436 §31. Counterexamples oo 1 Let V0= |~" -Bn. Then V0 contains no element u = (uij) with the n= 1 U property that, for each i, there is a coordinate u{j with \utj\ ^2. For if n 1 N |a I * uer-5„, then |wmfc| ^ V ^—^, with V|aJ^l, for m>N, so that in in i n i fact lii^l^l. We shall show that V0 can contain no ^(A)-neighbourhood of o. 00 Every bounded subset of A[£] is a subset of a set Q (c„J3J°, for n= 1 suitably chosen c„. Now (] (c„£„)° = r c„B„) = V°, so that we ob- tain a 3^-base of neighbourhoods of o in /x by taking the bipolars K°°, as V runs through the ^-neighbourhoods of o. We must therefore show oo that V0 can contain no V°°. Suppose that we are given V= \~ cnBn. «= i We denote by ttj the vector (utj) with utj=\ and uk = 0 for (fc,/) + (iJ). The element 2n + 1tnk lies in cnBn for sufficiently large kn. J\ n Consequently X™^ e„fcn belongs to P cnBn<^V. Thus the I 2 ' n= 1 oo weak limit 2 £ cM>kn belongs to K°°, but does not belong to V0, by the i remark made above. Thus we have shown that /x [Xh] is not bornological, and so /[£] is not distinguished. Using §19,9.(2), ip] can be considered as a closed linear sub- 00 space H of a topological product £[£] = TT E{ of (B)-spaces. Let us consider the topologies Xh(H) and %b(E) on E'/H1. We assert that 00 these topologies are different. For E,[_<£b(E)~\9 being © E'h is borno- i= 1 logical, by §28,4.(1), and so E'/H1, with the topology £*,(£), is also bornological, by §28,4.(2). On the other hand, as we have just seen, the space E'/H1 is not born^ogical under the topology Xb(H), as it is isomorphic to /x. This example, which is due to Kothe and Grothendieck (cf. Grothendieck [10]), was used by Amemiya [1] to give an example of a bornological (DF)-space whose strong bidual is not bornological.
Bibliography Alaoglu, L.: [1] Weak topologies of normed linear spaces. Ann. Math. 41, 252—267 (1940). Amemiya, I.: [1] Some examples of (F)- and (DF)-spaces. Proc. Japan Acad. 33, 169—171 (1957). Arens, R. F: [1] Duality in linear spaces. Duke Math. J. 14, 787—794 (1947). —, and J. L. Kelley: [1] Characterizations of the space of continuous functions over a compact Hausdorff space. Trans. Amer. Math. Soc. 62,499—508 (1947). Ballier, F: [1] Uber lineartopologische Algebren. J. reine angew. Math. 195, 42—75 (1956). Banach, S.: [1] Sur les operations dans les ensembles abstraits et leur application aux equations integrates. Fund. Math. 3, 133—181 (1922). 2] Sur les fonctionnelles lineaires, I, II. Studia Math. 1, 211—216, 223—239 1929). — [3] Theorie des operations lineaires (Monografje Matematyczne Tom I), Warszawa 1932. —, u. S. Mazur: [l]Zur Theorie derlinearen Dimension. Studia Math. 4,100—112 (1933). —, et H. Steinhaus: [1] Sur le principe de la condensation de singularites. Fund. Math. 9, 50—61 (1927). Bessaga, C, A. Pe^czynski and S. Rolewicz: [1] Some properties of the norm in F-spaces. Studia Math. 16, 183—192 (1957). Birkhoff, G.: [1] A note on topological groups. Compositio Math. 3, 427—430 (1936). — [2] Moore-Smith convergence in general topology. Ann. Math. 38. 39—56 (1937). — [3] Lattice theory, 3rd Ed., Amer. Math. Soc. Colloq. Publ. Vol. XXV, New York 1967. Bohnenblust, H. F, and A. Sobczyk: [1] Extensions of functionals on complex linear spaces. Bull. Amer. Math. Soc. 44, 91—93 (1938). Bourbaki, N.: [1] Sur les espaces de Banach. C. R. Acad. Sci. Paris 206,1701—1704 (1938). — [2] Sur certains espaces vectoriels topologiques. Ann. Inst. Fourier 2, 5—16 (1950). — [3] Elements de mathematique, Livre I, Theorie des ensembles, 4 Vols. Paris: Hermann & Cie. Act. Sci. et Ind. Nos. 1141, 1212, 1243, 1258 (1951—1958). — [4] Elements de mathematique, Livre II, Algebre, 6 Vols. Act. Sci. et Ind. Nos. 1144, 1032, 1044, 1102, 1179, 1261 (1947—1958). — [5] Elements des mathematique, Livre III, Topologie Generate, 6 Vols. Act. Sci. et Ind. Nos. 1142, 1143, 1029, 1045, 1084, 1196 (1947—1953). — [6] Elements de mathematique, Livre V, Espaces vectoriels topologiques, 2 Vols. Act. Sci. et Ind. Vols. 1189, 1229 (1953, 1955). — [7] Elements de mathematique, Livre VI, Integration, 2 Vols. Act. Sci. et Ind. Nos. 1175, 1244(1952, 1956). Bourgin, D. G.: [1] Linear topological spaces. Amer. J. Math. 65,637—659 (1943).
438 Bibliography Braconnier, J.: [1] Spectres d'espaces et de groupes topologiques. Portugaliae Math. 7, 93—111 (1948). Chillingworth, H. R.: [1] Generalised "dual" sequence spaces. Nederl. Akad. Wet. Proa, Ser. A 61, 307—315 (1958). Civin, P., and B. Yood: [1] Quasi-reflexive spaces. Proc. Amer. Math. Soc. 8, 906—911 (1957). Clarkson, J. A.: Uniformly convex spaces. Trans. Amer. Math. Soc. 40, 396—414 (1936). Collins, 'H. S.: [1] Completeness and compactness in linear topological spaces. Trans. Amer. Math. Soc. 79, 256—280 (1955). Cooke, R. G.: [1 ] Infinite matrices and sequence spaces. London: Macmillan & Co. 1950. — [2] Linear operators. London: Macmillan & Co. 1953. Cooper, J. L. B.: [1] Coordinated linear spaces. Proc. London Math. Soc, III. ser. 3, 305—327 (1953). Cudia, D. E: [1] Rotundity. Proc. Symp. Pure Math. VII, 73—97 (1963). — [2] The geometry of Banach spaces. Smoothness. Trans. Amer. Math. Soc. 110, 284—314(1964). Day, M. M.: [1] The spaces Lp with 0<p< 1. Bull. Amer. Math. Soc. 46, 816—823 (1940). — [2] Reflexive Banach spaces not isomorphic to uniformly convex spaces. Bull. Amer. Math. Soc. 47, 313—317 (1941). — [3] Some more uniformly convex spaces. Bull. Amer. Math. Soc. 47, 504—507 (1941). — [4] Uniform convexity III. Bull. Amer. Math. Soc. 49, 745—750 (1943). — [5] Uniform convexity in factor and conjugate spaces. Ann. Math. (2) 45, 375—385 (1944). — [6] Strict convexity and smoothness of normed spaces. Trans. Amer. Math. Soc. 78, 516—528 (1955). — [7] Every L-space is isomorphic to a strictly convex space. Proc. Amer. Math. Soc. 8, 415—417 (1957). — [8] Normed linear spaces. Ergebnisse d. Math., N. E, H. 21. Berlin-Gottingen- Heidelberg: Springer 1958. Dieudonne, J.: [1] Sur le theoreme de Hahn-Banach. Rev. Sci. 79, 642—643 (1941). — [2] La dualite dans les espaces vectoriels topologiques. Ann. Ecole Norm. 59, 107—139(1942). — [3] Sur la separation des ensembles convexes dans un espace de Banach. Rev. Sci. 81, 277—278 (1943). — [4] Matrices semi-finies et espaces localement lineairement compacts. J. reine angew. Math. 188, 162—166 (1950). — [5] Natural homomorphisms in Banach spaces. Proc. Amer. Math. Soc. 1, 54—59 (1950). — [6] Linearly compact spaces and double vector spaces over sfields. Amer. J. Math. 73, 13—19 (1951). — [7] Sur les espaces de Kothe. J. Analyse Math. 1, 81—115 (1951). — [8] Complex structures on real Banach spaces. Proc. Amer. Math. Soc. 3, 162—164 (1952). — [9] Sur les proprietes de permanence de certains espaces vectoriels topologiques. Ann. Soc. Polon. Math. 25, 50—55 (1952). — [10] Sur les sousespaces lineairement compacts. Bol. Soc. Math. Sao Paulo 6, 53—60 (1952). — [11] Sur un theoreme de Smulian. Arch. d. Math. 3, 436—440 (1953).
Bibliography 439 — [12] Recent developments in the theory of locally convex vector spaces. Bull. Amer. Math. Soc. 59, 495—512 (1953). — [13] Sur les espaces de Montel metrisables. C. R. Acad. Sci. Paris 238,194—195 (1954). — [14] Bounded sets in (F)-spaces. Proc. Amer. Math. Soc. 6, 729—731 (1955). — [15] Denumerability conditions in locally convex vector spaces. Proc. Amer. Math. Soc. 8, 367—372 (1957). —, et A. P. Gomes: [1] Sur certains espaces vectoriels topologiques. C. R. Acad. Sci. Paris 230, 1129—1130 (1950). —, et L. Schwartz: [1] La dualite dans les espaces (F) et (LF). Ann. Inst. Fourier 1, 61—101 (1950). Dixmier, J.: [1] Sur un theoreme de Banach. Duke Math. J. 15, 1057—1071 (1948). Donoghue, W. F, and K. T. Smith: [1] On the symmetry and bounded closure of locally convex spaces. Trans. Amer. Math. Soc. 73, 321—344 (1952). Dunford, N., and J. T. Schwartz: [1] Linear Operators, Part I: General Theory. New York: Interscience Publishers 1958. Eberlein, W. F: [1] Closure, convexity and linearity in Banach spaces. Ann. Math. (2) 47, 688—703 (1946). — [2] Weak compactness in Banach spaces, I. Proc. Nat. Acad. Sci. USA 33, 51—53 (1947). Eidelheit, M.: [1] Zur Theorie der konvexen Mengen in linearen normierten Raumen. Studia Math. 6, 104—111 (1936). Fan, K., and I. Glicksberg: [1] Fully convex normed linear spaces. Proc. Nat. Acad. Sci. USA 41, 947—953 (1955). — [2] Some geometric properties of the spheres in a normed linear space. Duke Math. J. 25, 553—568 (1958). Fichtenholz, G., et L. Kantorovitch : [1] Sur les operations dans l'espace des fonctions bornees. Studia Math. 5, 69—98 (1934). Fischer, H. R., and H. Gross: [1] Quadratic forms and linear topologies, I. Math. Ann. 157, 296—325 (1964). — [2] II. Math. Ann. 159, 285—308 (1965). — [3] III. Math. Ann. 160, 1—40 (1965). Fleischer, I.: [1] Locally symmetric spaces. Diss. Univ. of Chicago 1952. — [2] Sur les espaces normes non-archimediens. Nederl. Akad. Wet. Proc, Ser. A 57, 165—168 (1954). Freundlich-Smith, A.: [1] The Pontrjagin duality theorem in linear spaces. Ann. Math. 56,248—253(1952). Garling, D. J. H.: [1] Locally convex spaces with denumerable systems of weakly compact subsets. Proc. Cambridge Phil. Soc. 60, 813 -815 (1964). — [2] On topological sequence spaces. Proc. Cambridge Phil. Soc. 63, 977—1019 (1967). — [3] The /?- and y-duality of sequence spaces. Proc. Cambridge Phil. Soc. 63, 963—981 (1967). Goldstine, H. H.: [1] Weakly complete Banach spaces. Duke Math. J. 4, 125—131 (1938). Graves, L. M.: [1] On the completing of a Hausdorff space. Ann. Math. 38, 61—64 (1937). Grothendieck, A.: [1] Sur la completion du dual d'un espace vectoriel localement convexe. C. R. Acad. Sci. Paris 230, 605—606 (1950). — [2] Quelques resultats relatifs a la dualite dans les espaces (F). C. R. Acad. Sci. Paris 230, 1561—1563 (1950). — [3] Criteres generaux de compacite dans les espaces vectoriels localement con- vexes. Pathologie des espaces (LF). C. R. Acad. Sci. Paris 231, 940—942 (1950).
440 Bibliography — [4] Quelques resultats sur les espaces vectoriels topologiques. C. R. Acad. Sci. Paris 233, 839—841 (1951). — [5] Sur une notion de produit tensoriel topologique d'espaces vectoriels topologiques, et une classe remarquable d'espaces vectoriels liees a cette notion. C. R. Acad. Sci. Paris 233, 1556—1558 (1951). — [6] Criteres de compacite dans les espaces fonctionnels generaux. Amer. J. Math. 74, 168—186(1952). — [7] Sur les applications lineaires faiblement compactes d'espaces du type C(K). Canad. J. Math. 5, 129—173 (1953). — [8] Sur certains espaces de fonctions holomorphes, I, II. J. reine angew. Math. 192, 35—64, 77—95 (1953). — [9] Sur les espaces de solutions d'une classe generate d'equations aux derivees partielles. J. Analyse Math. 2, 243—280 (1953). — [10] Sur les espaces (F) et (DF). Summa Brasil. Math. 3, 57—123 (1954). — [11] Espaces vectoriels topologiques, Departamento de Matcmatica da Uni- versidade de Sao Paulo 1954. — [12] Resume des resultats essentiels dans la theorie des produits tensoriels topologiques et des espaces nucleaires. Ann. Inst. Fourier 4, 73—112 (1954). — [13] Produits tensoriels topologiques et espaces nucleaires. Mem. Amer. Math. Soc. No. 16 (1955). — [14] Une caracterisation vectorielle-metrique des espaces L1. Canad. J. Math. 7, 552—561 (1955). Hagemann, E.: [1] Das Reziprokentheorem in beliebigen linearen Koordinaten- raumen. Math. Ann. 114, 126—143 (1937). Hahn, H.: [1 ] Uber Folgen linearer Operationen. Mh. Math. Phys. 32,3—88 (1922). — [2] Uber lineare Gleichungssysteme in linearen Raumen. J. reine angew. Math. 157, 214—229 (1926). Halperin, I.: [1] Convex sets in linear topological spaces. Trans. Royal Soc. Canada 47, Ser. Ill, 1—6 (1953). — [2] Uniform convexity in function spaces. Duke Math. J. 21, 195—204 (1954). — [3] Reflexivity in the t-function spaces. Duke Math. J. 21, 205—208 (1954). Hammer, P. C: [1] Maximal convex sets. Duke Math. J. 22, 103—106 (1955). Hausdorff, F.: [1] Zur Theorie der linearen metrischen Raume. J. reine angew. Math. 167, 294—311 (1932). — [2] Set theory, 2nd Ed. New York: Chelsea Publishing Co. 1962. Hellinger, E., u. O. Toeplitz: [1] Integralgleichungen und Gleichungen mit un- endlich vielen Unbekannten. Enz. Math. Wiss. II C 13, 1335—1616 (1928). Helly, E.: [1] Uber lineare Funktionaloperationen. Sitzgsber. Akad. Wiss. Wien Math.-Nat. Kl. 121 II a, 265—297 (1912). — [2] Uber Systeme linearer Gleichungen mit unendlich vielen Unbekannten. Mh. Math. Phys. 31, 60—91 (1921). Hermes, H.: [1] Einfiihrungin die Verbandstheorie. Berlin-Gottingen-Heidelberg: Springer 1955. Hildebrandt, T. E.: [1] On bounded linear functional operations. Trans. Amer. Math. Soc. 36, 868—875 (1934). Hille, E., and R. S. Phillips: [1] Functional analysis and semigroups. Amer. Math. Soc. Colloq. Publ. Vol. XXXI, New York 1957. Hyers, D. H.: [1] Locally bounded linear topological spaces. Rev. Ci. Lima 41, 555—574(1939). — [2] Linear topological spaces. Bull. Amer. Math. Soc. 51, 1—21 (1945). Ingleton, A. W: [1] The Hahn-Banach theorem for non-Archimedean valued fields. Proc. Cambridge Phil. Soc. 48, 41—45 (1952).
Bibliography 441 Iyer, G.: [1] On the space of integral functions, I—III. J. Indian Math. Soc. (2) 12,13—30 (1948). — Quart. J. Math. (2) 1, 86—96 (1950). — Proc. Amer. Math. Soc. 3, 874—883 (1952). James, R. C: [1] Bases and reflexivity of Banach spaces. Ann. Math. 52, 518—527 (1950). — [2] A non-reflexive Banach space isometric with its second conjugate space. Proc. Nat. Acad. Sci. USA 37, 174—177 (1951). — [3] Projections in the space (m). Proc. Nat. Acad. Sci. USA 41, 899—902 (1955). — [4] Reflexivity and the supremum of linear functional. Ann. Math. 66,159—169 (1957). — [5] Characterizations of reflexivity. Studia Math. 23, 205—216 (1964). Kadison, R. V: [1] A representation theory for commutative topological algebra. Mem. Amer. Math. Soc. No. 7 (1951). Kakutani, S.: [1] Ein Beweis des Satzes von M. Eidelheit iiber konvexc Mengen. Proc. Imp. Acad. Tokyo 13, 93—94 (1937). — [2] Weak convergence in uniformly convex spaces. Tohoku Math. J. 45, 188—193 (1938). — [3] Weak topology and regularity of Banach spaces. Proc. Imp. Akad. Tokyo 15, 169—173 (1939). — [4] Some characterisations of Euclidean spaces. Jap. J. Math. 16,93—97 (1939). — [5] Weak topology, bicompact set and the principle of duality. Proc. Imp. Acad. Tokyo 16, 63—67 (1940). Kamke, E.: [1] Theory of sets. New York: Dover Publications Inc. 1950. Kaplan, S.: [1] Cartesian products of reals. Amer. J. Math. 74,936—954 (1952). Kelley, J. L.: [1] Note on a theorem of Krein and Milman. J. Osaka Inst. Sci. Tech., Parti, 3, 1—2(1951). — [2] General topology. New York: D. van Nostrand 1955. Klee, V. L.: [1] Some characterisations of reflexivity. Rev. Ci. Lima 52, 15—23 (1950). — [2] Convex sets in linaer spaces (I), II, III. Duke Math. J. 18, 443—466, 875—883 (1951); 20, 105—111 (1953). — [3] Invariant metrics in groups (solution of a problem of Banach). Proc. Amer. Math. Soc. 3, 484—487 (1952). — [4] Separation properties of convex cones. Proc. Amer. Math. Soc. 6, 313—318 (1955). — [5] The structure of semi-spaces. Math. Scand. 4, 54—56 (1956). — [6] Iteration of the "lin" Operation for convex sets. Math. Scand. 4, 231—238 (1956). — [7] Strict separation of convex sets. Proc. Amer. Math. Soc. 7, 735—737 (1956). — [8] Extremal structure of convex sets. Arch. d. Math. 8, 234—240 (1957). — [9] Extremal structure of convex sets, II. Math. Z. 69, 90—104 (1958). — [10] Some new results on smoothness and rotundity in normed linear spaces. Math. Ann. 139, 51—63 (1959). Kothe, G.: [1] Die konvergenzfreien Raume abzahlbarer Stufe. Math. Ann. Ill, 229—258 (1935). — [2] Die Teilraume eines linearen Koordinatenraumes. Math. Ann. 114,99—125 (1937). — [3] Losbarkeitsbedingungen fur Gleichungen mit unendlichvielen Unbekannten. J. reine angew. Math. 178, 193—213 (1938). — [4] Erweiterung von Linearfunktionen in linearen Raumen. Math. Ann. 116, 719—732 (1939). — [5] Die Qoutientenraume eines linearen vollkommenen Raumes. Math. Z. 51, 17—35 (1947).
442 Bibliography — [6] Die Stufenraume, eine einfache Klasse linearer vollkommener Raume. Math. Z. 51, 317—345 (1948). — [7] Eine axiomatische Kennzeichnung der linearen Raume vom Typus co. Math. Ann. 120, 634—649 (1949). — [8] Uber die Vollstandigkeit einer Klasse lokalkonvexer Raume. Math. Z. 52, 627—630 (1950). — [9] Uber zwei Satze von Banach. Math. Z. 53, 203—209 (1950). — [10] Neubegnindung der Theorie der vollkommenen Raume. Math. Nachr. 4, 70—80(1951). — [11 ] Die Randverteilungen analytischer Funktionen. Math. Z. 57, 13—33 (1952). — [12] Dualitat in der Funktionentheorie. J. reine angew. Math. 191, 30—49 (1953). — [13] Lineare Raume mit linearer Topologie. Proc. Int. Math. Congr. Amsterdam I, 236—237 (1954). — [14] Bericht uber neuere Entwicklungen in der Theorie der topologischen Vektorraume. Jber. DMV 59, 19—36 (1956). —, u. O. Toeplitz: [1] Theorie der halbfiniten unendlichen Matrizen. J. reine angew. Math. 165, 116—127 (1931). — [2] Lineare Raume mit unendlich vielen Koordinaten und Ringe unendlicher Matrizen. J. reine angew. Math. 171, 193—226 (1934). Kolmogoroff, A.: [1] Zur Normierbarkeit eines allgemeinen topologischen Rau- mes. Studia Math. 5, 29—33 (1934). Komatuzaki, H.: [1] Sur les projections dans certains espaces du type (B). Proc. Imp. Acad. Tokyo 16, 274—279 (1940). — [2] Une remarque sur les projections dans certains espaces du type (B). Proc. Imp. Acad. Tokyo 17, 238—240 (1941). Komura, Y.: [1] On linear topological spaces. Kumamoto J. Science, Series A, 5, Nr. 3, 148—157(1962). — [2] Some examples on linear topological spaces. Math. Ann. 153, 150—162 (1964). Komura, T, and Y.: [1] Sur les espaces parfaits de suites et leurs generalisations. J. Math. Soc. Japan 15, 319—338 (1963). Krein, M.: [1] Sur quelques questions de la geometrie des ensembles convexes situes dans un espace lineaire norme et complet. Dokl. Akad. Nauk SSSR, N. S. 14, 5—7 (1937). —, and D. Milman: [1] On extreme points of regularly convex sets. Studia Math. 9, 133—J38 (1940). —, and V. Smulian: [1] On regularly convex sets in the space conjugate to a Banach space. Ann. Math. 41, 556—583 (1940). Landsberg, M.: [1] Pseudonormen in der Theorie der linearen topologischen Raume. Math. Nachr. 14, 29—38 (1955). — [2] Lineare topologische Raume, die nicht lokalkonvex sind. Math. Z. 65, 104—112(1956). LaSalle, J. P.: [1] Pseudo-normed linear spaces. Duke Math. J. 8, 131—135 (1941). Leeuw, K. de, and W. Rudin: [1] Extreme points and extremum problems in Hv Pacific J. 8, 467—485 (1958). Lefschetz, S.: [1] Algebraic topology. Amer. Math. Soc. Colloq. Publ. Vol. XXVII, New York 1942. Leptin, H.: [1] Linearkompakte Moduln, I. Math. Z. 62, 241—267 (1955). — II. Math. Z. 66, 289—327 (1957). Lowig, H.: [1] Komplexe Euklidische Raume von beliebiger endlicher oder unendlicher Dimensionszahl. Acta Sci. Math. Szeged 7, 1—13 (1934).
Bibliography 443 Loomis, L. H.: [1] An introduction to abstract harmonic analysis. New York: D. van Nostrand 1953. Lorentz, G. G.: [1] Operations in linear metric spaces. Duke Math. J. 15,755—761 (1948). Lovaglia, A. R.: [1] Locally uniformly convex Banach spaces. Trans. Amer. Math. Soc. 78, 225—238 (1955). Mackey, G. W.: [1] On infinite dimensional linear spaces. Proc. Nat. Acad. Sci. USA 29, 216—221 (1943). — [2] On convex topological linear spaces. Proc. Nat. Acad. Sci. USA 29, 315—319 (1943). — [3] Equivalence of a problem in measure theory to a problem in the theory of vector lattices. Bull. Amer. Math. Soc. 50, 719—722 (1944). — [4] On infinite-dimensional linear spaces. Trans. Amer. Math. Soc. 57, 155—207 (1945). — [5] On convex topological linear spaces. Trans. Amer. Math. Soc. 60, 519—537 (1946). Matthews, G.: [1] Generalised rings of infinite matrices. Nederl. Akad. Wet. Proc, Ser. A 61, 298—306 (1958). Mazur, S.: [1] Uber die kleinste konvexe Menge, die eine gegebene kompakte Menge enthalt. Studia Math. 2, 7—9 (1930). — [2] Uber konvexe Mengen in linearen normierten Raumen. Studia Math. 4, 70—84 (1933). — [3] Uber schwache Konvergenz in den Raumen Lp. Studia Math. 4, 128—133 (1933). —, u. W. Orlicz: [1] Uber Folgen linearer Operationen. Studia Math. 4, 152—157 (1933). — [2] Sur les espaces metriques lineaires I, II. Studia Math. 10, 184—208 (1948); 13, 137—179(1953). McShane, E. J.: [1] Linear functionals on certain Banach spaces. Proc. Amer. Math. Soc. 1, 402—408 (1950). Michael, E. A.: [1] Locally multiplicatively-convex topological algebras. Mem. Amer. Math. Soc. No. 11 (1952). Milman, D. P.: [1] On some criteria for the regularity of spaces of the type (B). Dokl. Akad. Nauk SSSR, N. S. 20, 243—246 (1938). — [2] Characteristics of extremal points of regularly convex spaces. Dokl. Akad. Nauk SSSR, N. S. 57, 119—122 (1947) [Russian]. —, and M. A. Rutman: [1] On a more precise theorem about the completeness of the system of extreme points of a regularly convex set. Dokl. Akad. Nauk SSSR, N. S. 60, 25—27 (1948) [Russian]. Monna, A. F.: [1] Sur les espaces lineaires normes, I—VI. Nederl. Akad. Wet. Proc. 49,1045—1055,1056—1062,1134—1141,1142— 1152(1946); 51,197—210 (1948); 52, 151—160(1949). — [2] Espaces lineaires a une infinite denombrable de coordonnees. Nederl. Akad. Wet. Proc. 53, 1548—1559 (1950). Murray, F. J.: [1] On complementary manifolds and projections in spaces Lp and lp. Trans. Amer. Math. Soc. 41, 138—152 (1937). — [2] Quasi-complements and closed projections in reflexive Banach spaces. Trans. Amer. Math. Soc. 58, 77—95 (1945). Nachbin, L.: [1] Espagos vetoriais topologicos. Rio de Janeiro: Livraria Boffoni 1948. — [2] A characterisation of the normed vector ordered spaces of continuous functions over a compact space. Amer. J. Math. 71, 701—705 (1949).
444 Bibliography — [3] A theorem of the Hahn-Banach type for linear transformations. Trans. Amer. Math. Soc. 68, 28—46 (1950). — [4] Topological vector spaces of continuous functions. Proc. Nat. Acad. Sci. USA 40, 471—474(1954). Naimark, M. A.: [1] Normed rings. Groningen: P. NoordhoffN. V. 1959. Nakano, H.: [1] Topology and linear topological spaces. Tokyo: Maruzen Co. 1951. Natanson, I. P.: [1] Theory of functions of a real variable. 2 Vols. New York: Frederick Ungar Publishing Co. 1955, 1960. Neumann, J. von: [1] On complete topological linear spaces. Trans. Amer. Math. Soc. 37, 1—20 (1935). Nikodym, O. M.: [1] On transfinite iterations of the weak linear closure of convex sets in linear spaces, A. B. Rend. Circ. Mat. Palermo, II. ser. 2, 85—105 (1953); 3, 5—75 (1954). Ornstein, D.: [1] Dual vector spaces. Ann. Math. 69, 520—534 (1959). Pettis, B. J.: [1] A note on regular Banach spaces. Bull. Amer. Math. Soc. 44, 420—428 (1938). — [2] A proof that every uniformly convex space is reflexive. Duke Math. J. 5, 249—253 (1939). Phillips, R. S.: [1] On linear transformations. Trans. Amer. Math. Soc. 48, 516—541 (1940). — [2] On weakly compact subsets of a Banach space. Amer. J. Math. 65,108—136 (1943). Pietsch, A.: [1] Verallgemeinerte vollkommene Folgenraume. Schriftenreihe d. Institute d. Math. Deutsche Akad. Wiss. Heft 12, Akademie-Verlag Berlin (1962). Ptak, V.: [1] On complete topological linear spaces. Cehosl. Mat. Z. 3 (78), 301—364 (1953) [Russian]. — [2] Weak compactness in convex topological linear space. Cehosl. Mat. Z. 4, (79), 175—186(1954). — [3] On a theorem of W. F. Eberlein. Studia Math. 14, 276—284 (1954). — [4] Completeness and the open mapping theorem. Bull. Soc. math. France 86, 41—74(1958). — [5] A combinatorial theorem on systems of inequalities and its application to analysis. Cehosl. Mat. Z. 9 (84), 629—630 (1959). — [6] A combinatorial lemma on the existence of convex means and its application to weak compactness. Proc. Symp. Pure Math. VII, 437—450 (1963). Rai'kow, D. A.: [1] Wpolne neprerywnyje spektry lokalno wypuklych prostranstw. Trudy Mockowskogo Mat. Obschestwa 7, 413—438 (1958). Riesz, F.: [1] Sur les operations fonctionnelles lineaires. C. R. Acad. Sci. Paris 149, 974—977 (1909). — [2] Les systemes d'equations lineaires a une infinite d'inconnues. Paris: Gauthier- Villars 1913. — [3] Uber lineare Funktionalgleichungen. Acta Math. 41, 71—98 (1918). —, and B. Sz. Nagy: Functional Analysis. London and Glasgow: Blackie & Son 1956. Ritzdorff, F.: [1] Das Tragheitsgesetz der quadratischen Formen mit halbfiniter Koeffizientenmatrix. Math. Z. 44, 23—54 (1938). Roberts, G. T.: [1] The bounded-weak topology and completeness in vector spaces. Proc. Cambridge Phil. Soc. 49, 183—189 (1953). Robertson, W.: [1] Contributions to the general theory of linear topological spaces. Thesis, Cambridge 1954.
Bibliography 445 — [2] Completions of topological vector spaces. Proc. London Math. Soc III. ser. 8, 242—257 (1958). Rolewicz, S.: [1] On a certain class of linear metric spaces. Bull. Acad. Polon Sci., CI. Ill 5, 471—473 (1957). Ruston, F.: [1] A note on convexity in Banach spaces. Proc. Cambridge Phil Soc. 45, 157—159(1949). — [2] A short proof of a theorem on reflexive spaces. Proc. Cambridge Phil. Soc 45, 674 (1949). — [3] Conjugate Banach spaces. Proc. Cambridge Phil. Soc. 53, 576—580 (1957). Schatten, R.: [1] The space of completely continuous operators on a Hilbert space. Math. Ann. 134, 47—49 (1957). Schauder, J.: [1] Zur Theorie stetiger Abbildungen in Funktionalraumen. Math. Z. 26, 47—65, 417—431 (1927). — [2] Uber die Umkehrung linearer stetiger Funktionaloperationen. Studia Math. 2,1—6(1930). Schubert, H.: [1] Topology. London: MacDonald. 1968. Schwartz, L.: [1] Theorie des distributions. New Ed. Paris: Hermann & Co. 1966. — [2] Espaces de fonctions differentiates a valeurs vectorielles. J. Analyse Math. 4, 88—148 (1955). Sebastiao e Silva, J.: [1] As funcoes analiticas e a analise funcional. Port. Math. 9, 1—130(1950). — [2] Sobre a topologia dos espacos funcionais analiticos. Rev. Fac. Cienc. Lisboa, II. ser. A 1, 23—102 (1950). — [3] Sui fundamenti della teoria dei funzionali analitici. Port. Math. 12, 1—46 (1953). — [4] Su certe classi di spazi localmente convessi importanti per le applicazioni. Rend. Mat. Roma, V. ser. 14, 388—410 (1955). Shirota, T.: [1] On locally convex vector spaces of continuous functions. Proc. Jap. Acad. 30, 294—298 (1954). Sierpinski, W.: [1] Sur les ensembles complets d'un espace (D). Fund. Math. 11, 203—205 (1928). Silva Dias, C. L. da : [1 ] Espacos vectoriais topologicos e sua aplicacao nos espacos funcionais analiticos. Bol. Soc. Mat. Sao Paulo 5, 1—58 (1950). Smulian, V.: [1] Sur les ensembles regulierement fermes et faiblement compacts dans 1'espace du type (B). C. R. URSS 18, 405—407 (1938). — [2] On the principle of inclusion in the space of type (B). Mat. Sbornik, N. S. 5, 327—328 (1939) [Russian]. — [3] On some geometrical properties of the unit sphere in the space of the type (B). Mat. Sbornik, N. S. 6, 77—94 (1939) [Russian]. — [4] Sur la derivabilite de la norme dans l'espace de Banach. Dokl. Akad. Nauk SSSR, N. S. 27, 643—648 (1940). — [5] Uber lineare topologische Raume. Mat. Sbornik, N. S. 7, 425^48 (1940). — [6] Sur la structure de la sphere unitaire dans l'espace de Banach. Mat. Sbornik, N. S. 9, 545—561 (1941). — [7] Sur les espaces lineaires topologiques, II. Mat. Sbornik, N. S. 9, 727—730 (1941). — [8] Sur les ensembles compacts et faiblement compacts dans l'espace du type (B). Mat. Sbornik, N. S. 12, 91—95 (1943). Sobczyk, A.: [1] Projections in Minkowski and Banach spaces. Duke math. J. 8,78—106(1941). — [2] Projections of the space (m) on its subspace (c0). Bull. Amer. Math. Soc. 47, 938—947 (1941). 29 Kothe, Topological Vector Spaces I
446 Bibliography — [3] On the extension of linear transformations. Trans. Amer. Math. Soc. 55, 153—169(1944). Straszewicz, S.: [1] Uber exponierte Punkte abgeschlossener Punktmengen. Fund. Math. 24, 139—143 (1935). Suchomlinoff, G. A.: [1] Uber Fortsetzung von linearen Funktionalen in linearen komplexen Raumen und linearen Quotientenraumen. Mat. Sbornik, N. S. 3, 353—358 (1938) [Russian]. Takenouchi, O.: [1] Sur les espaces lineaires localement convexes. Math. J. Okayama Univ. 2, 57—84 (1952). Taylor, A. E.: [1] Banach spaces of functions analytic in the unit circle. Studia Math. 11, 145—170 (1950); 12, 25—50 (1951). — [2] Introduction to Functional Analysis. New York. John Wiley & Sons 1958. Tillmann, H. G.: [1] Randverteilungen analytischer Funktionen und Distribu- tionen. Math. Z. 59, 61—83 (1953). — [2] Dualitat in der Potentialtheorie. Port. Math. 13, 55—86 (1954). — [3] Dualitat in der Funktionentheorie auf Riemannschen Flachen. J. reine angew. Math. 195, 76—101 (1956). — [4] Die Fortsetzung analytischer Funktionale. Hamb. Abh. 21, 139—193 (1957). Toeplitz, O.: [1] Uber die Auflosung unendlich vieler linearer Gleichungen mit unendlich vielen Unbekannten. Rend. Circ. Mat. Palermo 28, 88—96 (1909). — [2] Die linearen vollkommenen Raume der Funktionentheorie. Comment. Math. Helv. 23, 222—242 (1949). Tukey, J. W.: [1] Some notes on the separation of convex sets. Port. Math. 3, 95—102 (1942). Tychonoff, A.: [1] Ein Fixpunktsatz. Math. Ann. Ill, 767—776 (1935). Ulam, S.: [1] Zur MaBtheorie in der allgemeinen Mengenlehre. Fund. Math. 16, 140—150(1930). Ulm, H.: [1] Elementarteilertheorie unendlicher Matrizen. Math. Ann. 114, 493—505 (1937). Vilenkin, N. Y.: [1] Vector spaces over topological fields. Mat. Sbornik 74, 195—208 (1953) [Russian]. Warner, S., and A. Blair: [1] On symmetry in convex topological vector spaces. Proc. Amer. Math. Soc. 6, 301—304 (1955). Wehausen, J. V.: [1] Transformations in linear topological spaces. Duke math. J. 4, 157—169(1938). Weil, A.: [1] Sur les espaces a structure uniforme et sur la topologie generate. Act. Sci. etInd.No.551(1938). — [2] L'integration dans les groupes topologiques et ses applications. Act. Sci. et Ind. No. 869 (1940). Yosida, K., and E. Hewitt: [1] Finitely additive measures. Trans. Amer. Math. Soc. 72, 46—66 (1952). Zaanen, A. C.: [1] Linear Analysis. Groningen: P. Noordhoff 1953. Zelinsky, D.: [1] Linearly compact modules and rings. Amer. J. Math. 75, 79—90 (1953). Zeller, K.: [1] FK-Raume in der Funktionentheorie I. Math. Z. 58, 288—305 (1953). — II. Math. Z. 58, 414—435 (1953). — [2] Theorie der Limitierungsverfahren. Erg. d. Math., N. F., H. 15. Berlin- Gottingen-Heidelberg: Springer 1958. Zorn, M.: [1] A remark on method in transfinite algebra. Bull. Amer. Math. Soc. 41, 667—670 (1935).
Author and Subject Index Ka-compact 18,315 a-dual 405 absolute bipolar 246, 249 — polar 245 absolutely convex 160, 173 ff, 203 about a point 175 cover 173ff., 240 ff., 325 — p-convex 160 cover 160 absorb 301 absorbent 145 accumulation point 4 adherent point of a filter 12 of a net 10 adjoint mapping 73, 197, 237 Alaoglu, L. 248 Alaoglu-Bourbaki theorem 248, 264 AlexandrofT, P. 21 Alexandroff s theorem 21 algebra 59 algebraic basis 51,194 — boundary 177 point 177 — complement 51 — conjugate 69 — dimension 53, 75 — dual 69 ff., 88, 97, 101 ff, 247 — hull 177 — interior point 177 — kernel 177 — point of smoothness 345 algebraically closed 177, 193, 194 convex a-body, cf. convex a-body half space 179ff. — isomorphic 53 — open 177 — open half space 179ff Allen, H. S. 423 almost constant sequence 89 Amemiya, I. 384, 404, 405, 436 anti-isomorphic 65 antisymmetric 9 Arens, R. F. 260, 335 associated bornological space 380 asymptote 341 automorphism 60 BV(I) 425 j5-dual 423 (B)-space 126, 250, 252, 273, 280, 283, 303,304,315,335,336,389,401, 431, cf. also (F)-space Baire, R. 27 Baire's theorem 27 Ballier, F. 121 Banach, S. 165,168,170,189, 259, 272, 350, 431 Banach algebra 130 — space 126 Banach's theorem 169 Banach-Dieudonne theorem 252, 254, 272 Banach-Mackey theorem 252, 254 Banach-Schauder theorem 166 Banach-Steinhaus theorem 169 Banach-Stone theorem 334 barrel 257 barrelled 257,261,297,305,306,367 ff, 369,371,372,380,434 base of a uniform space 30 — of neighborhoods 3 basis of open sets 1 Bessaga, C. 166 bidual 129, 196, 298, 300 ff, 388 — space, cf. bidual bilinear form 78 — functional 78 — mapping 78, 171 ff. bipolar 246 Birkhoff,G. 11 Bohnenblust, H. F. 192 Boolean algebra 58 bornological 379ff, 387, 388, 399 ff, 400,403,419,434,436 29*
448 Author and Subject Index boundary 4 — point 4 bounded 24, 152, 248 ff., 254, 403 ff. — above 9 — below 9 — closure 386 ff., 401 boundedly closed 386ff., 400 Bourbaki, N. 1, 12, 20, 76, 121, 172, 173,186, 188, 211, 233, 260, 312, 332, 357,366,427,431,434 Bourbaki's theorem 172 Bourgin, D.G. 159, 162 Braconnier, J. 233 C[/] 138, 197ff.,260,430 C(K), C(R) 138, 250, 323ff., 334, 335, 343, 350 c 131 c0 131,302,333,335,343,425 canonical mapping of E onto E/H 60 — representation in the narrow sense 67 in the wide sense 67 Cantor, G. 25 Cartesian product 8 Cauchy filter 32,210 — net 32 — sequence 25 character 309 characteristic function 41 Chillingworth, H. R. 423 circled 146 — cover 146, 174,241 Civin, P. 304 Clarkson, J. A. 353, 357 close of order N 29 closed 321, 322 — absolutely convex cover 175, cf. also absolutely convex cover set, cf. absolutely convex — ball of radius r 24 — convex cover 175, cf. also convex cover set, cf. convex — graph theorem 167 — linear subspace, cf. linear subspace — mapping 6 — set 1 closure 4 — point 4, 312 coarser filter 12 — topology 5 coarser uniformity 30 co-dimension 55 co-echelon space 419, 433 p-th order 420 cofinal 10 — subnet 10 Collins, H.S. 269,271 column 63 column-finite matrix 63 compact 16ff., 154, 241 ff., 279, 313ff., 326,331,336,340,385,415 compactum 26 compatible 82, 145 — locally convex topology 236, 245, 254, 261 — linear topology 236 complement of the image space 60 complementary subspaces 51, 95, cf. also topological complement complemented lattice 58 complete 25, 32, 165, 210, 231, 269 ff., 402, 435, cf. also quasi-complete, sequentially complete — in itself 252 — lattice 57, 85 — metric space 25, 42, — metrizable vector space 166 ff., 172 completely regular 45, 47 completion of a Hausdorff uniform space 33 — of a linearly topologized space 115, 149 — of a locally convex space 208, 248, 261,269 — of a metric space 25 — of a topological vector space 148, 158 complex hyperplane 180 — linear functional 179 — locally convex space 273 ff. — vector space 49 cone 183ff, 195, 337 ff. — generated by a set 184 conjugate space 86, 128 connected 5, 152 continuous 6 — basis 101 — bilinear mapping 171 continuous dimension 101,102 — linear functional 156,158 — linear mapping 98, 129, 148, 166, 167,237,262,291,297,333,381
Author and Subject Index 449 continuous projection, cf. projection co-nullity 67 converge 10, 12 convergent sequence of matrices 107 convex 160,173 ft, 186 ff, 194,244,273, 322, 336, 337, 343 ff. — a-body 180—194 — algebraic body, cf. convex a-body convex-compact (weakly) 316 convex cover 173ff, 240ff., 245, 321, 322,325,331 — function 181 — 2-body 180, 182, 187, 188, 193, 342 Cooke, R. G. 423 Cooper, J. L. B. 424 coordinate space 405 countability axioms 19, 20 countable at infinity 22 — degree 120, 370 countably compact 19, 310ff, 315 316 Cudia, D. 366 Day, M. 157, 317, 318, 360, 361, 363, 366 defect 55, 67, 103 dense 4 density zero 369 (DF)-space 396 ff, 401 ff, 434 diagonal 29 — transform 408 diameter of a set 24 diametrically opposite cone 183 Dieudonne, J. 48,85,113,121,272,274, 311, 318, 354, 369, 370, 371, 372, 384, 387,404,419,421,424,432 dimension, cf. algebraic —, continuous — direct product 76 — sum 54, 57 directed set 9 discrete topology 4, 82, 83 — uniformity 31 distance between two points 23 — between two sets 24 distinguished 306 ff, 399, 400, 435 distributive lattice 58 Dixmier, J. 304, 336, 426 Donoghue, W. F. 392 dual 86, 128, 275 ff, 298, cf. also iterated duals — pair 85, 234 — space, cf. dual dual system 70 dually isomorphic lattices 57 £"' 308,426 Eberlein, W. F. 313 Eberlein's theorem 313, 317, 326, 366, 415 Abounded 251,254 echelon space 419, 433 of p-th order 420 Eidelheit, M. 187 embedding 60 endomorphism 59 equicontinuous 168, 172, 258, 259 — bilinear mapping 171 ff. equivalent base of neighbourhoods 3 — defining system 216, 227 — filter bases 12 — in the narrow sense 67 — in the wide sense 68 — linear mappings 67 — norms 125 — system of equations 105 — uniform spaces 30 Erdos, P. 75 Erdos and Kaplansky's theorem 75 essential supremum 142 essentially bounded 142 Euclidean space 23 everywhere dense 4 exposed point 337 — ray 341 extension of a linear functional 70, 86, 188 ff, 233 — theorem, cf. preceding entry exterior point 4 extreme point 330, 337, 338, 340, 346 — ray 337 Fan, K. 366 Fantappie, L. 373 Fichtenholz, G. 425, 427 filter 11 filter-base 12 filter corresponding to a net 11 finer filter 12 — topology 5 — uniformity 30 finite coordinate space 53 — dimensional topological vector spaces 151 ff. — vector 53
450 Author and Subject Index finitely additive 424 first axiom of countability 19 — category 28 Fischer, E. 142 Fischer, H. R. 121 Fleischer, I. 121 (F)-norm 163ff. Frechet, M,. 164 Frechet derivative 364 Freundlich-Smith, A. 309 (F)-space 164, 205, 225, 265, 273, 278, 303, 306, 309, 315, 318, 371, 388, 389, 393, 399, 400 ff, 431 (FK)-space 424 (FM)-space 369, 373, 433 function 6 fundamental sequence of bounded sets 392 — set 132,237 — system of neighbourhoods 3 Garling, D. J. H. 372, 423 Gateaux derivative 349 gauge 46 Glicksberg, I. 366 Gomes, A. P. 371,421 graph 167 greatest lower bound 9 of a collection of functions 41 Gross, H. 121 Grothendieck, A. 115, 225, 265, 266, 269,270, 310, 313, 318, 324, 326, 378, 392, 396, 398, 403, 404, 405, 434,436 Grothendieck's theorem 269 HB{&) 138ff.,352ff. //(©) 372ff, 432 H(r) 421 H(D) 375 Hagemann, E. 121,431 Hahn, H. 189 Hahn-Banach theorem 188ff, 196 half-space 179 Hammer, P. C. 186 Hausdorff, F. 252, 260 Hausdorff separation axiom 3 Hausdorff space 3 Helly,E. 189 Henriques, P. 270 Hermes, H. 10 Hewitt, E. 425 Hilbert cube 29 Hilbert parallelotope 29 Hilbert space 23 Hille,E. 130 Holder's inequality 135, 140 homeomorphism 2 homogeneous 103 homomorphism 59 hull topology 215 Hyers, D.H. 159,162 hypercone 184 hyperplane 56, 157, 180 image-filter 13 image set 6 — space 6, 60 indicatrix 373 induced topology 4 inductive limit 219 Ingleton, W. A. 121 inhomogeneous 103 injection 60 integrable 139 interchangeable double limits 326 interior 4 — point 4 internal point 176ff. intersection, a a b 57 — of topologies 5 inverse 6,61 — image 6 inverse-image filter 13 — space 60 inversely directed 9 invertible 61,66 involution 428 isolated point 4 isometric 23 isomorphic algebras 65 — dual pairs 85 — lattices 57 — uniform spaces 30 — vector spaces 53 iterated duals 304, cf. also bidual, E James, R. C 300,366,431 Kadison, R. V. 335 Kakutani, S. 187,431 Kamke, E. 9, 10 Kantorovitch, L. V. 425, 427 Kaplansky, I. 75
Author and Subject Index 451 Kaplansky's theorem 312 Kelly, J. L. 1,305 kernel 60,226,292 — topology 226, 292 Klee, V. L. 165,173, 178, 194, 268, 322, 337, 339, 341, 347 Klee's theorems 319 ff. Kolmogroff, A. 145,160 Komatuzaki, H. 431 Komura, T. 419, 423 Komura, Y. 222, 268, 369, 388, 404, 419, 423 Kothe, G. 48, 113, 120, 121, 378, 415, 421,423,432,434,436 Krein, M. 325 Krein's theorem 325, 326, 336 Krein-Milman theorem 331, 336ff., 339 Kronecker product 81 L1 333, 335, 343, 350, 357, cf. also E L°° 142, 333, 343, 347, 352, 357, 425 U 139ff., 156, 343, 351, 355 ff., 431 Lp0<p<\, 156, 160, 195 Z1 131,280, 281,282, 311, 333, 343, 347, 351, 363, 406, 426, cf. also lp \\ 283, 404, 424, cf. also /J /°° 130,283,333,343, 347,352,363,406, 424, 426, cf. also /" I? 283, 363, 424, cf. also /J lp 136, 195, 333, 343, 351, 352, 355, 358, 407, 430 /J 137,359,431 /"(£„) 359,360 lpn 137,429 Landsberg, M. 160, 162 lattice 57 (LB)-space 223 ff., 434 least upper bound 9 of a collection of functions 41 Lebesgue, H. 252 Lebesgue's theorem 324 Leeuw, K. de 353 Lefschetz, S. 1, 3, 48, 82, 95, 109, 115, 233 Lefschetz's theorem 109, 115 left inverse 61,66 — reciprocal 61 length of a vector 106 Leptin, H. 121 (LF)-space 223 ff., 384 limes inferior 39 — superior 39 limit 10, 12 linear combination 50 — equation 103 ff. — form 69 — functional 69 — group 60 — manifold 50 — mapping 59 — matrix ring 65 — metric space 164ff. — span 50 — space 48 — strong topology, cf. Xlb — subspace 50,148,207,275,279, 282, 384, 401, 424 ff. — system 85 — topology 82 — transformation 59 — weak topology, cf. Xls linearly bounded 113, 116 — compact 95, 108 — dependent 50 — independent 50 — precompact 116 — topologized space 83 local convergence 382 locally bounded linear functional 379 — bounded linear mapping 381 — bounded topological vector space 159 ff. — compact 20, 42, 155, 250, 338, 340, 341, 343, 345 — continuous 383 — convex direct sum 211 ff., 283 ff., 296, 299, 303, 370 hull 215ff., 292ff., 368, 381, 383, 402 kernel 226, 292 topological vector space 202, 208, 232, 250, 258 — holomorphic 139, 375 — linearly compact 108 — uniformly convex 366 Loomis, L. H. 130 Lovaglia, A. R. 366 lower bound 9 — limit 39 — semi-continuous 40, 258 2R(K), <m(R) 138, 324, 333, 339, 343, 347
452 Author and Subject Index Mackey, G. W. 48, 85, 260, 380, 384, 387, 392 Mackey's theorem 254 Mackey-Arens theorem 98, 260 Mackey convergence 382 — topology, cf. Xk Mackey-Ulam theorem 392 mapping , 6 matrix 63 — product 64 Matthews, G. 423 maximal 9 — linear matrix ring 65 Mazur,S. 188,259,347,351,366 McShane E. J. 357 meagre 28, 166 measure 138 Meray, C. 25 metric 23 — space 23 metrizable 43,45, 163 ff., 170, 205, 228, 245, 251, 263, 265, 272, 273, 301, 311, 380, 393 ff., 400, 401, 403, cf. also (F)-space Milman, D. P. 332, 336, 354 Milman's theorem 332, 341 Milman-Rutman theorem 336, 337 minimal 9 — supporting facet 336 ff. Minkowski, H. 123 Minkowski functional 159, 180 Minkowski's inequality 135,140 modular lattice 58 module of concavity 161 — of convexity 353 Monna, R. F. 121 monomorphism 60 monotonic decreasing net 38 — increasing net 38 Montel space, cf. (M)-space Montel's theorem 373 (M)-space 369,421,434 Murray, F. J. 428 Nachbin, L. 121,380 Naimark, M.A. 130 Natanson, I. P. 198, 199 natural embedding 276 — homomorphism 93, 275 — topology 300 nearest point 343 nearest-point mapping 344 neighbourhood 2 — filter 12 — of a set 4 net 10 — corresponding to a filter 12 Neumann, J. von 145 Neumer, W. 75 Nikodym, O. 178 norm 123 — isomorphic 125 normable 150, 160,393 normal 22 — cover 407, 417 — form 113 — sequence space 405 — set 408 — topology of a sequence space 407 normalized 200 normed algebra 130 — space 123,263,279,337,342 ff, 403, cf. also metrizable space nowhere dense 4 nullity 66, 103 null-space 60 co, md 56,70,75,119,122,151,155,243, 248,268,287,391,406,431,432 a><p 119, 120,121,431 a)(p@(pa) 304,370 one-point compactification 21 open ball of radius r 24 — mapping 6, cf. also topological homomorphism — neighbourhood 374 — set 1 Ornstein, D. 431 orthogonal 70 — space 70 orthogonally closed 71, 85, 120, 236 tyA 8 <p, <pd 53,70,76,122, 214, 268,287,308, 370,406,431 <pa) 119, 120ff,431 <pa)@a)(p 120 ff, 240, 296, 304, 370 \jj 109, 122 parallel 50 parallelotope 8 partial continuity 8 partially compact (weakly) 317 — ordered 9 p-convex 160
Author and Subject Index 453 Pefczynski, A. 166 perfect 406 Phillips, R. S. 130, 325, 427 Pietsch, A. 423 p-norm 160 p-normable 160 point at infinity 22 — measure 324 — of smoothness 345 ff., 349, 350 — of support 193 pointed cone 183 pointwise convergence 323 polar 245 — reflexive 308 — semi-reflexive 308 — topology 266 Pontryagin, L. 310 Pontryagin duality theorem 310 positive measure 334, 339 — vector 407 positive-homogeneous 180, 188 precompact 26, 36, 153, 155, 266, 273, 385 product-filter 14 product of uniform spaces 37 — of vector spaces 56 projection 8, 60, 428, 431, cf. also topological complement — of a filter 14 projective limit 229 ff. proper cone 183 pseudo-compact 315 Ptak,V. 271,316,318,330 Ptak's lemma 316 Ptak's theorem 326 quadratic form 122 quasi-barrelled 301, 367 ff, 379, 388, 396, 399, 400 quasi-closed 296 quasi-closure 296 quasi-complete 210, 278, 295 ff, 299, 305, 309, 313, 318, 319 ff, 325, 402 quasi-completion 297, 306 quasi-norm 159 quasi-reflexive 304 quotient space 50, 90, 99, 120, 127, 149 ff, 207, 275, 279, 296, 304, 368, 383,401,434,435 topology 90, 127, 149 P, P" 5 pco g Raikow, D. A. 233 rank 66,78,103 real hyperplane 180, 186ff, 243, 320, 322 — locally convex space 208, 243, 273 ff, 328, 344 — vector space 49 reciprocal 61 reduced form 290 reflexive 9, 129, 302 ff, 308, 309, 315, 354, 360, 369, 372, 389, 396, 400, 434 — dual pair 304 ff. regular 15 relatively compact 17 restriction of a filter 12 Riesz, F. 142,200,201 Riesz's theorem 200 Riesz-Fischer theorem 142 right inverse 61, 66 — module 121 — vector space 121 Ritzdorff,K. 122 Robertson, W. 157,211 Rolewicz, S. 161,162,166 rotund 342 row 63 Rudin, W. 353 Ruston, F 336 Rutman, M. A. 336 a-additive measure 426 saturated 255 — cover 256 Schatten, R. 336 Schauder, J. 166 Schwartz, L. 311,372 Sebastiao e Silva, J. 233, 378 second axiom of countability 20 — category 28 section 412 sectional subspace 410 semi-finite 109, 112 semi-norm 124,203 semi-ordered 9 semi-reflexive 298,304,305,306,307 ff, 315,319,320,322,355 separable 26, 126, 128, 259, 260, 271, 273, 280, 283,362, 370, 371, 398, 400, 401, 403 separated 3, 187
454 Author and Subject Index separately continuous 171 ff. — equicontinuous 171 ff. separation theorem 187ff., 243, 322 for compact convex sets 243 sequence spaces 405 sequentially closed 273, 313 — compact 19, 310ff., 313 — complete 89, 210, 295, 296 — continuous 11,271, 383 set-theoretic product 8 Shirota, T. 380 shortest distance 343 ff. shortness of a vector 106 Sierpinski, W. 165 Silva Dias, C. L. da 378 similar automorphisms 121 ff simply ordered 9 "sliding hump" 252, 281 small of order N 32 Smith, K. T. 392 smooth 346 ff. smoothly normable 361 ff, 363 Smulian, V. 311,366 Smulian's theorem 311 Sobczyk, A. 192,428,430 solution of equations 103 sphere 24 square matrix 65 Steinhaus, H. 169 steps 419 Stone, M. H. 334 Straszewicz, S. 337 strict closure point 296 — inductive limit 222, 370 — (LB)-space 223 — (LF)-space 223, 312, 319 strictly convex 342, 346 ff — normable 361 ff — separated 187, 243 strong bidual 300 — derivative 364 ff. — dual 115,257,306,388 — topology, cf. Zb stronger topology 5 strongly bounded 251,252, 254 strongly differentiable 364 — inaccessible 392 — reflexive 115, 118 — semi-reflexive 115 sub-additive 180, 188 sub-base 12 sub-basis 1 subspace, cf. linear subspace — of an involution 428 sum of mappings 59 — of matrices 64 — of subspaces 54, 152, 154, 322 summable 139 support manifold 330 supporting facet 336 — hyperplane 193 ff, 244, 245, 320, 349 symmetric mapping 122 — sequence space 409 — vicinity 30 Sz.-Nagy, B. 201 Zb 256, 257, 263, 277, 285, 286, 300, 370, 385, 400, 436 Zb> 262, 263, 283, 286, 398 Zc 263, 264, 273, 278, 385, 389 Z 385 ff Zr 267ff,269,271 Zk 98,260, 263, 277, 278,282,285, 286, 293, 302, 325, 385 Zlb 115 Zlf 267ff,269,271 £tt 97 Zls 86ff,238 £TO 255,266 Zn 300 r 266,271 Z 323 ZPS 234,238,248,276,277,285,286, 354 Z* 380 Takenouchi, O. 233 tangent hyperplane 346 tensor product 76 theorem of Alaoglu-Bourbaki 248, 264 — of Alexandroff 21 — ofBaire 27 — of Banach-Dieudonne 252, 254, 272 — of Banach-Mackey 252, 254 — of Banach-Schauder 166 — of Banach-Steinhaus 169 — of Banch-Stone 334 — ofbipolars 246 — ofBourbaki 172 — of Eberlein 313, 317, 326, 366, 415 — of Erdos and Kaplansky 75 — of Grothendieck 269
Author and Subject Index 455 theorem of Hahn-Banach 188ff., 196 — ofKaplansky 312 — ofKlee 319ff. — of Krein 325, 326, 336 — of Krein-Milman 331, 336 ff, 339 — ofLebesgue 324 — ofLefschetz 109, 115 — ofMackey 254 — of Mackey-Arens 98, 260 — of Mackey-Ulam 392 — of Milman 332, 341 — of Milman-Rutman 336, 337 — of Montel 373 — ofPontryagin 310 — ofPtak 326 — ofRiesz 200 — of Riesz-Fischer 142 — ofSmulian 311 — ofTychonoff 18,96 — of Urysohn 44, 45 Tillmann, H. G. 378 Toeplitz, O. 48,113,252,378,421,423, 432 topological complement 92, 95, 99, 109,121, 156,158, 168, 238, 239, 240, 424ff, 426 — direct sum 84, 117, 214, 288, 308 — group (abelian) 309 — homomorphism 91, 150, 166 — inductive limit 220, 224, 289, 290, 403 — isomorphism 84, 91, 125, 150 — linear space 145 — monomorphism 91, 150 — product 7, 8, 117, 127, 149 ff., 207, 283 ff, 296, 299, 303, 368, 370, 384, 389 — projective limit 230 ff, 290,294,300 — space 1 — sum 100 — vector space 83, 145 ff. topologically isomorphic 84, 125 topology 1 — of a uniform space 30 — of pointwise convergence, cf. 2p — of precompact convergence, cf. £c — of uniform convergence on 90? 255 total 132,237,255 totally bounded 26, 36 — disconnected 6, 83 — ordered 9 transitive 9,227 translation-invariant 124, 147, 164 transposed equation 103 — matrix 74 triangular matrix 106 trivial topology 5 truncated cone 183 T2 -space 3 Tychonoff, A. 45,151 Tychonoff space 45 Tychonoff s theorem 18,96 Ulam, S. 392 Ulm, H. 122 ultrafilter 14 uniform norm topology 130 — space 29, 47 uniformity 29, 386 uniformizable 43, 147 uniformly continuous 24, 32, 265 — convex 353, 360, 365, 366 — equicontinuous 168 — normable 361 ff. — smooth 363 ff. — strongly differentiate 364 union, a v b 57 — of topologies 5 upper bound 9 — limit 39 — semi-continuous 40 Urysohn, P. 42 Urysohn's embedding theorem 45 Urysohn's extension theorem 44 Urysohn's lemma 42 variation 425 vector 48 — space 48 vertex 183 vicinities 29 Vilenkin, N. Y. 121 weak derivative 349 ff. — dual 235 — neighbourhood 85 — topology, cf. Xs weaker topology 5 weakly complete 89, 248 — continuous 237
456 Author and Subject Index weakly convex-compact 316 Yood, B. 304 — differentiable 349 Yosida, K. 425 — partially compact 317 — precompact 248 Zelinsky, D. 121 Weil, A. 29,233 Zeller, K. 424 well-ordered 9 Zorn, M. 9 Zorn's Lemma 9, 10 Typesetting and printing: Zechnersche Buchdruckcrei, Speyer
Die Grundlehren der mathematischen Wissenschaften in Einzeldarstellungen mit besonderer Beriicksichtigung der Anwendungsgebiete 2. Knopp: Theorie und Anwendung der unendlichen Reihen. DM 48,—; US $ 12.00 3. Hurwitz: Vorlesungen uber allgemeine Funktionentheorie und elliptische Funk- tionen. DM 49,—; US $ 12.25 4. Madelung: Die mathematischen Hilfsmittel des Physikers. DM 49,70; US $ 12.45 10. Schouten: Ricci-Calculus. DM 58,60; US $ 14.65 14. Klein: Elementarmathematik vom hoheren Standpunkt aus. 1. Band: Arithmetik. Algebra. Analysis. DM 24,— ; US $ 6.00 15. Klein: Elementarmathematik vom hoheren Standpunkt aus. 2. Band: Geometrie. DM24,— ; US$6.00 16. Klein: Elementarmathematik vom hoheren Standpunkt aus. 3. Band: Prazisions- und Approximationsmathematik. DM 19,80; US $ 4.95 19. Polya/Szego: Aufgaben und Lehrsatze aus der Analysis I: Reihen, Integralrechnung, Funktionentheorie. DM 34,— ; US $ 8.50 20. Polya/Szego: Aufgaben und Lehrsatze aus der Analysis II: Funktionentheorie, Nullstellen, Polynome, Determinanten, Zahlentheorie. DM 38,—; US $ 9.50 22. Klein: Vorlesungen uber hohere Geometrie. DM 28,— ; US $ 7.00 26. Klein: Vorlesungen uber nicht-euklidische Geometrie. DM 24,—; US $ 6.00 27. Hilbert/Ackermann: Grundzuge der theoretischen Logik. DM 38,—; US $ 9.50 30. Lichtenstein: Grundlagen der Hydromechanik. DM 38,— ; US $ 9.50 31. Kellogg: Foundations of Potential Theory. DM 32,-; US $ 8.00 32. Reidemeister: Vorlesungen uber Grundlagen der Geometrie. DM 18,—; US $ 4.50 38. Neumann: Mathematische Grundlagen der Quantenmechanik. DM 28,—; US $ 7.00 40. Hilbert/Bernays: Grundlagen der Mathematik I. DM 68,— ; US $ 17.00 43. Vorlesungen uber Geschichte der antiken mathematischen Wissenschaften. Band 1: Neugebauer: Vorgriechische Mathematik. Nachdruck in Vorbereitung. 50. Hilbert/Bernays: Grundlagen der Mathematik II. DM 68, -; US $ 17.00 52. Magnus/Oberhettinger/Soni: Formulas and Theorems for the Special Functions of Mathematical Physics. DM 66,— ; US $ 16.50 57. Hamel: Theoretische Mechanik. DM 84,— ; US $ 21.00 58. Blaschke/Reichardt: Einfuhrung in die Differentialgeometrie. DM 24,—; US $ 6.00 59. Hasse: Vorlesungen uber Zahlentheorie. DM 69,— ; US $ 17.25 60. Collatz: The Numerical Treatment of Differential Equations. DM 78,— ; US $ 19.50 61. Maak: Fastperiodische Funktionen. DM 38,— ; US $ 9.50 62. Sauer: Anfangswertprobleme bei partiellen Differentialgleichungen. DM 41,—; US$ 10.25 64. Nevanlinna: Uniformisierung. DM 49,50; US $ 12.40 66. Bieberbach: Theorie der gewohnlichen Differentialgleichungen. DM 58,50; USS 14.65 68. Aumann: Reelle Funktionen. DM 59,60; US $ 14.90 69. Schmidt: Mathematische Gesetze der Logik I. DM 79,— ; US $ 19.75 71. Meixner/Schafke: Mathieusche Funktionen und Spharoidfunktionen mit Anwen- dungen auf physikalische und technische Probleme. DM 52,60; US $ 13.15 73. Hermes: Einfuhrung in die Verbandstheorie. DM 46,—; US $ 11.50
74. Boerner: Darstellungen von Gruppen. DM 58,— ; US $ 14.50 75. Rado/Reichelderfer: Continuous Transformations in Analysis, with an Introduction to Algebraic Topology. DM 59,60; US $ 14.90 76. Tricomi: Vorlesungen uber Orthogonalreihen. DM 37,60; US $ 9.40 77. Behnke/Sommer: Theorie der analytischen Funktionen einer komplexen Verander- lichen. DM 79,— ; US $ 19.75 78. Lorenzen: Einfuhrung in die operative Logik und Mathematik. DM 54,—; US,$ 13.50 80. Pickert: Projektive Ebenen. DM 48,60; US $ 12.15 81. Schneider: Einfuhrung in die transzendenten Zahlen. DM 24,80; US $ 6.20 82. Specht: Gruppentheorie. DM 69,60; US $ 17.40 84. Conforto: Abelsche Funktionen und algebraische Geometric DM 41,80; US $ 10.45 86. Richter: Wahrscheinlichkeitstheorie. DM 68,— ; US $ 17.00 87. van der Waerden: Mathematische Statistik. DM 49,60; US $ 12.40 88. Miiller: Grundprobleme der mathematischen Theorie elektromagnetischer Schwin- gungen. DM 52,80; US $ 13.20 89. Pfluger: Theorie der Riemannschen Flachen. DM 39,20; US $ 9.80 90. Oberhettinger: Tabellen zur Fourier Transformation. DM 39,50; US $ 9.90 91. Prachar: Primzahlverteilung. DM 58,—; US $ 14.50 92. Rehbock: Darstellende Geometric DM 29,— ; US $ 7.25 93. Hadwiger: Vorlesungen uber Inhalt, Oberflache und Isoperimetrie. DM 49,80; US$ 12.45 94. Funk: Variationsrechnung und ihre Anwendung in Physik und Technik. DM 98,—; US $ 24.50 95. Maeda: Kontinuierliche Geometrien. DM 39,— ; US $ 9.75 97. Greub: Lineare Algebra. DM 39,20; US $ 9.80 98. Saxer: Versicherungsmathematik. 2. Teil. DM 48,60; US $ 12.15 99. Cassels: An Introduction to the Geometry of Numbers. DM 69,— ; US $ 17.25 100, Koppenfels/Stallmann: Praxis der konformen Abbildung. DM 69,—; US $ 17.25 101. Rund: The Differential Geometry of Finsler Spaces. DM 59,60; US $ 14.90 103. Schutte: Beweistheoric DM 48,— ; US $ 12.00 104. Chung: Markov Chains with Stationary Transition Probabilities. DM 56,—; US $ 14.00 105. Rinow: Die innere Geometrie der metrischen Raume. DM 83,—; US $ 20.75 106. Scholz/Hasenjaeger: Grundziige der mathematischen Logik. DM 98,—; US $ 24.50 107. Kothe: Topologische Lineare Raume I. DM 78,— ; US $ 19.50 108. Dynkin: Die Grundlagen der Theorie der Markoffschen Prozesse. DM 33,80; US $ 8.45 109. Hermes: Aufzahlbarkeit, Entscheidbarkeit, Berechenbarkeit. DM 49,80; US $ 12.45 110. Dinghas: Vorlesungen uber Funktionentheorie. DM 69,—; US $ 17.25 111. Lions: Equations differentielles operationnelles et problemes aux limites. DM 64,—; US$ 16.00 112. Morgenstern/Szabo: Vorlesungen uber theoretische Mechanik. DM 69,—; US$ 17.25 113. Meschkowski: Hilbertsche Raume mit Kernfunktion. DM 58,— ; US $ 14.50 114. MacLane: Homology. DM 62,— ; US $ 15.50 115. Hewitt/Ross: Abstract Harmonic Analysis. Vol. 1: Structure of Topological Groups. Integration Theory. Group Representations. DM 76,—; US $ 19.00 116. Hormander: Linear Partial Differential Operators. DM 42,— ; US $ 10.50
117. O'Meara: Introduction to Quadratic Forms. DM 48,—; US $ 12.00 118. Schafke: Einfiihrung in die Theorie der speziellen Funktionen der mathematischen Physik. DM49,40; US $ 12.35 119. Harris: The Theory of Branching Processes. DM 36,—; US $ 9.00 120. Collatz: Funktionalanalysis und numerische Mathematik. DM 58,—; US $ 14.50 121. Dynkin: Markov Processes. DM 96,— ; US $ 24.00 122. 123. Yosida: Functional Analysis. DM 66,— ; US $ 16.50 124. Morgenstern: Einfiihrung in die Wahrscheinlichkeitsrechnung und mathematische Statistik. DM 38- ; US $ 9.50 125. Ito/McKean: Diffusion Processes and Their Sample Paths. DM 58,— ; US $ 14.50 126. Letho/Virtanen: Quasikonforme Abbildungen. DM 38,— ; US $ 9.50 127. Hermes: Enumerability, Decidability, Computability. DM 39,— ; US $ 9.75 128. Braun/Koecher: Jordan-Algebren. DM 48,— ; US $ 12.00 129. Nikodym: The Mathematical Apparatus for Quantum-Theories. DM 144,—; US $ 36.00 130. Morrey: Multiple Integrals in the Calculus of Variations. DM 78,— ; US $ 19.50 131. Hirzebruch: Topological Methods in Algebraic Geometry. DM 38,— ; US $ 9.50 132. Kato: Perturbation Theory for Linear Operators. DM 79,20; US $ 19.80 133. Haupt/Kunneth: Geometrische Ordnungen. DM 68,— ; US $ 17.00 134. Huppert: Endliche Gruppen I. DM 156,— ; US $ 39.00 135. Handbook for Automatic Computation. Vol. 1/Part a: Rutishauser: Description of Algol 60. DM 58,— ; US $ 14.50 136. Greub: Multilinear Algebra. DM 32,— ; US $ 8.00 137. Handbook for Automatic Computation. Vol. 1/Part b: Grau/Hill/Langmaack: Translation of Algol 60. DM 64,— ; US $ 16.00 138. Hahn: Stability of Motion. DM 72,— ; US $ 18.00 139. Mathematische Hilfsmittel des Ingenieurs. Herausgeber: Sauer/Szabo. 1. Teil. DM88,—; US $22.00 140. Mathematische Hilfsmittel des Ingenieurs. Herausgeber: Sauer/Szabo. 2. Teil. DM 136,— ; US$34.00 141. Mathematische Hilfsmittel des Ingenieurs. Herausgeber: Sauer/Szabo. 3. Teil. DM98,—; US $24.50 142. Mathematische Hilfsmittel des Ingenieurs. Herausgeber: Sauer/Szabo. 4. Teil. In Vorbereitung 143. Schur/Grunsky: Vorlesungen iiber Invariantentheorie. DM 32,—; US $ 8.00 144. Weil: Basic Number Theory. DM 48,— ; US $ 12.00 145. Butzer/Berens: Semi-Groups of Operators and Approximation. DM 56,—; US $ 14.00 146. Treves: Locally Convex Spaces and Linear Partial Differential Equations. DM 36,—; US $ 9.00 147. Lamotke: Semisimpliziale algebraische Topologie. DM 48,— ; US $ 12.00 148. Chandrasekharan: Introduction to Analytic Number Theory. DM 28,— ; US $ 7.00 149. Sario/Oikawa: Capacity Functions. DM 96,— ; US $ 24.00 150. Iosifescu/Theodorescu: Random Processes and Learning. DM 68,—; US $ 17.00 151. Mandl: Analytical Treatment of One-dimensional Markov Processes. DM 36,—; US $ 9.00 152. Hewitt/Ross: Abstract Harmonic Analysis. Vol. II. In preparation
153. Federer: Geometric Measure Theory. DM 118,— ; US $ 29.50 154. Singer: Bases in Banach Spaces I. In preparation 155. Miiller: Foundations of the Mathematical Theory of Electromagnetic Waves. DM 58,— ; US $ 14.50 156. van der Waerden: Mathematical Statistics. DM 68,— ; US $ 17.00 157. Prohorov/Rozanov: Probability Theory. DM 68,— ; US $ 17.00 159. Kothe: Topological Vector Spaces I. DM 78,— ; US $ 19.50 160. Agrest/Maksimov: Theory of Incomplete Cylindrical Functions and their Applications. In preparation 161. Bhatia/Szego: Stability Theory of Dynamical Systems. In preparation 162. Nevanlinna: Analytic Functions. Approx. DM 68.—; US $ 17.00 163. Stoer/Witzgall: Convexity and Optimization in Finite Dimensions I. In preparation 164. Sario/Nakai: Classification Theory of Riemann Surfaces. In preparation