Mathematics and Its Applications
V. V. Filippov Basic Topological Structures of Ordinary Differential Equations
Kluwer Academic Publishers
Mathematics and Its Applications
Managing Editor: M. HAZEWINKEL Centre for Mathematics and Computer Science, Amsterdam, The Netherlands
Volume 432
Basic Topological Structures of Ordinary Differential Equations by
V. V. Filippov Department of Mathematics, Moscow State University, Moscow, Russia
KLUWER ACADEMIC PUBLISHERS DORDRECHT / BOSTON / LONDON
A c.I.P. Catalogue record for this book is available from the Library of Congress.
ISBN 0-7923-4951-2
Published by Kluwer Academic Publishers, P.O. Box 17,3300 AA Dordrecht, The Netherlands. Sold and distributed in the U.S.A. and Canada by Kluwer Academic Publishers, 101 Philip Drive, Norwell, MA 02061, U.S.A. In all other countries, sold and distributed by Kluwer Academic Publishers, P.O. Box 322, 3300 AH Dordrecht, The Netherlands.
This is a completely revised and updated translation of the original Russian works Solution Spaces of Ordinary Differential Equations, Moscow University Publishers, @1993 (in Russian), and Ordinary Differential Equations: Topological Approach to the Theory (to appear in Russian), both by V. V. Filippov.
Printed on acid-free paper
All Rights Reserved @1998 Kluwer Academic Publishers No part of the material protected by this copyright notice may be reproduced or utilized in any form or by any means, electronic or mechanical, including photocopying, recording or by any information storage and retrieval system, without written permission from the copyright owner Printed in the Netherlands.
Table of Contents
1
PREFACE
xi
TOPOLOGICAL AND METRIC SPACES Sets, mappings. 1 Topological spaces. 2 Some notions related to a topology. 3 Metric spaces. 4 Limits. 5 Compactness and completeness. 6 Continuous mappings .. 7 Zorn's lemma and well ordered sets. 8
1 1 8 12 14 17
23 27 33
2 SOME PROPERTIES OF TOPOLOGICAL, METRIC AND EUCLIDEAN SPACES 37 1 Tychonoff topology of products. . . . . . . . . . . . . . 37 2 Continuous bijections. Homeomorphisms. Metrizability. 42 3 Some properties of a metric. .. . . . . . . . . . . . . . 46 4 Some properties of Euclidean, locally compact, and normed 48 spaces . . . . . . . . . . . . . . . . . . 5 54 Remarks on multi-valued mappings. 6 Connectedness. . . . . . . . . . . . . 56 7 Plane regions. . . . . . . . . . . . . 61 8 Degree of a mapping of a circle into itself.. 63 3 SPACES OF MAPPINGS AND SPACES OF COMPACT SUBSETS 71 1 Metric and norm of uniform convergence. 71 Compact open topology. 2 73 3 Vietoris topology. . . . . . 77 4 Hausdorff metric. . . . . . 79 5 Space of partial mappings. 83 Compactness in the space of partial mappings. 6 87 7 Space Cs(M). . . . . . . . . . . . . . . . . . . 89
vi 8
Alexander's lemma and its consequences.
4 DERIVATION AND INTEGRATION 1 Measure of a set on the line. 2 Measurable functions. 3 Derivation of non-decreasing functions. 4 Derivative with respect to a set. 5 Absolutely continuous functions. 6 Derivation of absolutely continuous functions. 7 Lebesgue integral. 8 Density points and approximate derivatives. 9 Generalized absolutely continuous functions and the Denjoy integral. 10 Mean Value theorem. 11 Change of variable III an integral and derivation of a composite function. 12 Measurable multi-valued mappings. 13 Multi-valued mappings defined on products.
92
95 95 104 108 116 118 123 125 138 139 144 145 152 156
5
WEAK TOPOLOGY ON THE SPACE Ll AND DERIVATION OF CONVERGENT SEQUENCES 163 1 Continuous linear functionals. . . . 163 2 Hahn-Banach theorem. . . . . . . . 166 Convex sets and linear functionals. . 169 3 171 4 Weakly convergent sequences. 5 Spaces Ll and Loo. . . . . . . 173 Common representation of a linear functional on the space 6 L 1• . . . • • . . • • . . . . . . . . . . 175 7 Space of measurable sets. . . . . . . . . . . . . . . . . . .. 179 8 Weak compactness in the space L 1 . . . . . . . . . . . . . . 180 9 Derivation of a convergent equi-absolutely continuous sequence of functions. . . . . . . . . . . . . . . . . . . . . . . . . . .. 189
6
BASIC PROPERTIES OF SOLUTION SPACES Ordinary differential equations.. . . . . . . . 1 Solutions of differential inclusion y' E F(t, y). 2 The inequality Ily'(t)11 ~ cp(t). . . . . . . . . 3 Some properties of solution spaces.. . . . . . 4 Cauchy problem. Existence and uniqueness theorems. 5 Maximal extensions of solutions. . . . . . . . . . . . . 6 Continuity of the dependence of solutions on initial values. 7
191 191 192 195 197 199 204 207
Vll
7 CONVERGENT SEQUENCES OF SOLUTION 209 SPACES 1 Continuity of the dependence of solutions of equations on parameters of right hand sides. . . . . . . 209 2 Properties of convergent space sequences. 215 3 Equicontinuity condition. . . . . . . . . . 219 Estimates of set location. . . . . . . . . . 232 4 Estimates of the upper limit of a space sequence. . 235 5 6 Additional remarks. . . . . . 242 7 R(U) as a topological space. .. . . . . . . . 243 8 PEANO, CARATHEODORY AND DAVY 247 CONDITIONS 1 Kneser condition. . . . . . . . . . . 247 2 Caratheodory and Davy conditions. 253 3 Localization principle. . . . . . . . . 262 4 Absolutely continuous solutions. . . 264 5 Continuously differentiable solutions. Peano and Picard conditions.. . . . . . . . . . . . . . . . . . 266 269 6 General approach to existence theorems. 7 Majorants of right hand sides. . . . . . . 271 9
COMPARISON THEOREM 1 Existence of upper solutions in the scalar case. 2 Scorza-Dragoni property. . ..... . 3 Comparison theorem: scalar case. . . . . 4 Comparison theorem: n-dimensional case. 5 Localness of the property in question. . .
277 278 280 282 284 288
10 CHANGES OF VARIABLES, MORPHISMS AND MAXIMAL EXTENSIONS 291 1 Change of variables: general remarks. . . . . . . 291 2 Change of variables in equations and inclusions. 295 304 3 Maximal extensions. Some generalizations. 4 Convergent space sequences again. . . . . . . . . 309 5 Morphisms. . . . . . . . . . . . . . . . . . . . . . 313 6 One more remark about the continuity of the dependence of solutions on the right hand side. 316 7 Uniqueness theorem. 321
VIll
11 SOME METHODS OF INVESTIGATION OF EQUATIONS 327 1 Linear equations (systems). . . . . . . . . . . . . . . . . . 327 . . . . . . . . . . . . . . . . . 334 2 Linear equation of order n. 3 Modification of the right hand side. Equations with discontinuous 335 right hand sides. . . . . . . . . . . . . . . . . 4 Simplest singularities. Existence of solutions. 339 5 Collage of spaces. . . . . . . . . . . . . 342 6 Simplest singularities. Approximations. 344 7 Optimization. 350 8 Control. . . . . . . . . . . . . . . . . . . 352 12 EQUATIONS AND INCLUSIONS WITH COMPLICATED DISCONTINUITIES IN THE SPACE VARIABLES 357 1 Sufficient conditions for equi-absolute continuity. . . . . .. 358 Continuity of the dependence of solutions on the right hand 2 side. First step. . . . . . . . . . . . . . . . . . . . . . . . .. 368 3 Existence theorems. First step. . . . . . . . . . . . . . . . 373 4 Continuity of the dependence of solutions on the right hand side. Second step: complication of singularities. 378 5 Existence theorem. Second step. . . . . . . . . . . . . . .. 384 13 EQUATIONS AND INCLUSIONS OF SECOND ORDER. CAUCHY PROBLEM THEORY 389 1 Cauchy problem theory for differential inclusions of second order. . . . . . . . . . . . . . . . . . . . . . . . . . 389 394 2 Relatively weakly compact families of majorants. . Equation x" + f(x)x' + h(t, x) = o. ........ 402 3 14 EQUATIONS AND INCLUSIONS OF SECOND ORDER. PERIODIC SOLUTIONS, DIRICHLET PROBLEM 409 1 First remark on existence of periodic solutions. . . 409 2 Second remark on existence of periodic solutions. 415 3 Upper and lower solutions of the inclusion x" E F(t, x). 431 15 BEHAVIOR OF SOLUTIONS 1 Autonomous and asymptotically autonomous spaces. 2 Limit sets. .. . . . . . . . . . . . . . . . . . . Some geometric properties of solution spaces. 3 4 Periodic spaces. . . . . . . . . . . . . . . . . . 5 Limit passages in the space Rc(U).. . . . . . . 6 First approximation. Asymptotic stability of a stationary point. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
441 441 446 449 452 453 457
IX
7 8
Asymptotic estimates. . . . . . . . . . . Behavior of a solution space as Ilyll ---? 00.
16 TWO-DIMENSIONAL SYSTEMS 1 Existence of a stationary point. ..... . 2 Indices. . .................. . 3 Calculation of index by the right hand side. 4 Poincare-Bendixson theorem. . . . . . . . . 5 Neighborhoods of stationary points in the plane.
460 466 469
469 480 486 493 500
REFERENCES
507
INDEX
510
NOTATION
515
PREFACE
The aim of this book is a detailed study of topological effects related to continuity of the dependence of solutions on initial values and parameters. This allows us to develop cheaply a theory which deals easily with equations having singularities and with equations with multivalued right hand sides (differential inclusions). An explicit description of corresponding topological structures expands the theory in the case of equations with continuous right hand sides also. In reality, this is a new science where Ordinary Differential Equations, General Topology, Integration theory and Functional Analysis meet. In what concerns equations with discontinuities and differential inclusions, we do not restrict the consideration to the Cauchy problem, but we show how to develop an advanced theory whose volume is commensurable with the volume of the existing theory of Ordinary Differential Equations. The level of the account rises in the book step by step from second year student to working scientist. If you open any book on the general theory of Ordinary Differential Equations you can note the following property of its structure. One or two first chapters are devoted to proofs of a few basic theorems (about existence and uniqueness of solutions, about the continuous dependence of solutions on initial values etc.). Then authors pass on to some further account repeating in new theorems the list of hypotheses of the basic theorems mentioned (for instance, the continuity of the right-hand side), but in general they repeat the hypotheses (the continuity, the Caratheodory conditions, etc.) for the unique purpose to receive the possibility of referring to the basic theorems. This observation excites a desire to give an account of the theory independently of the contents of the introductory chapters in a way in which we may take basic properties of solutions as axioms of the new theory. We make some steps in this direction in the main part of this book. On the other hand a considerable part of our account is related just to the contents of the introductory chapters of the classical theory of Ordinary
xu Differential Equations. We try to develop powerful tools of investigation there, where usually is only a short list of preliminary propositions. This gives the possibility of an easy expansion of some parts of the theory to equations with various singularities and differential inclusions. One of our aims in this book is to show how they need word results in order to apply them to equations with very wide range of singularities. A comparison of proofs of results of our book with proofs of their classical prototypes allows one to affirm that such steps are not always trivial. Here we do not set ourselves the target of solving any concrete known problems in the theory of Ordinary Differential Equations. The subject of our investigation is the internal structure of the theory. Of course, we meet in this way results which are interesting not only in the framework of the theory in question, but they are here a secondary product. In general we realize here a known thesis that such domains of mathematics as topology, algebra, etc., are devoted not to solving any applied problems but to restructuring from time to time common building of Mathematics. For a long time my range of interests in mathematics was limited to the General Topology. In the beginning of the 1980s I began to realise that such a purely topological phenomenon as continuity of the dependence of solutions of ordinary differential equations on initial values and parameters had never been the subject of an investigation by my colleagues, experts in topological structures, but that it was nevertheless in need of such an investigation. This is reflected in our usual use of words. In practice one never meets the term 'continuity of the dependence of solutions on initial values and parameters'. People always speak about 'the continuous dependence of solutions on initial values and parameters', which corresponds to avoiding the analysis of the situation. Nevertheless under a close consideration by an expert in General Topology the necessity of the analysis seems to be obvious. And so I made the decision. It became clear that corresponding passages to the limit may be applied to the investigation of singularities, and to proofs of some assertions such as existence theorems. Mostly important properties of solution sets were distinguished. Their key position in the theory of Ordinary Differential Equations was evident; however, I have not found any signs of the existence of a theory in which these properties, or any similar ones, have played the roles of axioms, although attempts to follow this direction did took place (see [Za], [KI], [Yo]). None of these attempts led to any complete result. Discussions with experts did not help me in this search. This stimulated further investigations, and their result is presented in this book. Our aim here is to show that the topological structures found (which correspond approximately to topological relations mentioned in Theorem
Preface.
xiii
1.2.4 of the Hartman's book [Hal) are not an accidental observation, but they pierce the entire theory of Ordinary Differential Equations. In the old classical account of the theory they are hidden very deeply. Pulling them up to the surface we lay bare interior relations of the theory, we simplify their assimilation. It already happens at the level of equations with the continuous right hand sides, but the new theory works just as easily with differential inclusions and with differential equations having discontinuous right hand sides. Topological considerations of this book may be jointed well with the homological approach to boundary value problems, see [Go] and [Fvv5-6]. We cannot give an account here of the preliminary material needed and we leave aside corresponding new results which allow us to conjugate the simplicity of the method of translation along trajectories with the power of continuation and Leray-Schauder methods, see [Fvv4]. After some articles in journals the first connected account of purely topological aspects of the new theory was published in Chapter IX of the book [FF]. In preparing of the book [Fvvl] the author had a quite well developed theory. This book includes the material of the books [Fvvl] and [Fvv2] and reflects results of further investigations. As a monograph the book sums up scientific investigations. At the same time it is written as a textbook, which introduces a second or third year student in mathematics to a circle of new ideas. Almost at the beginning of the work on the new approach it had become clear that the material of some common mathematical courses is not expounded in a way which suits us (this may be the cause of why the theory of Ordinary Differential Equations is always developed without remarking on its internal topological basis). Therefore we here pay great attention to introductory material which is expounded in just the way that will satisfy the needs of our theory. In part the material of the book has been included by the author in his Courses in the Faculty of Mechanics and Mathematics of the Moscow State University. In the first three chapters we introduce the topological tools needed. In Chapter 4 we develop Integration theory adequate to our level of the theory of Ordinary Differential Equations (such as the Peano and Picard conditions corresponding to the Riemann integral, the Caratheodory and Davy conditions corresponding to the Lebesgue integral). Amongst other notions we consider here generalized absolutely continuous functions and the Denjoy integral. We meet the necessity of their application, for instance, when we investigate the equation y' = f(t, y) with a right hand side f which
XIV
is continuous everywhere except on a singleton where we do not know the type of the singularity. In Chapter 5 we give an account of some material from Functional Analysis which is important for the new theory. In particular, we give here a corresponding elementary version of the Dunford-Pettis theorem. In the remaining chapters we go on to give an account of the new approach to the theory of Ordinary Differential Equations. Notice a particularity of our account. We do not restrict the study to abstract topological considerations, but we show how to investigate right hand sides in order to apply topological observation to particular equations and concrete problems. Interior references in the book are made in the following manner. 'Theorem 5.4.1' means the theorem 4.1 which is proved in the section 4 of the chapter 5. In references inside a chapter the first number is omitted. '§5.4' denotes Section 4 of Chapter 5. '§4' denotes Section 4 of the chapter where the reference is made. V.V.Filippov.
CHAPTER 1
TOPOLOGICAL AND METRIC SPACES
It is natural to assume that any reader, who is familiar with any aspect of the theory of Ordinary Differential Equations, has already mastered basic topological notions as 'open set', 'closure', 'closed set', 'continuous mapping' ('continuous function'), etc .. But aiming here to simplify our later account, we recall in this chapter some of these notions and related theorems. We omit here proofs of theorems which may be assumed known by any reader who has a minimal training, except theorems which are based on arguments that illustrate well notions under consideration and relations between them. In order that the omissions of proofs do not make real difficulties in comprehension of the material we give references to books [AI] and [En], but these cases concern well known facts and the reader may use with -about the same success other sources as textbooks on Mathematical or Functional Analysis. Lastly, he can consider the restoration of omitted proofs as an exercise within his powers (after some basic teaching).
We do not assume known some notions, and there our account is rather complete. We keep here to manners of reasoning which are usual in introductory chapters of Higher Mathematics, although our theorems deal with more general situations. We have the concrete purpose of preparing the reader for the study of the topological structure of solution sets of ordinary differential equations, and we stop in our account whenever proving a more general result can lead us away from our chosen path. 1. Sets, mappings We use the word' set' as an equivalent of the words 'totality', 'collection'. Members of 'the totality' are called elements of the corresponding set. For instance, we may speak about the set of (all) points of the real line or of the plane. A set may be described by presenting a full list of its elements. Usually one writes the list inside braces: {2,4,5,6} denotes the set which contains the numbers 2, 4, 5 and 6 (and does not contain any other elements).
2
CHAPTER 1
The notation sEA (as the notation A 3 s) means that s is an element of the set A (one says also that s belongs to A or that A contains s). The notation s (j. A (as the notation A ;i s) means that s is not an element of the set A (s does not belong to A or A does not contain s). The notation A ~ B (as the notation B ;2 A) means that the set A is a part of the set B (one says also that A is a subset of the set B or that A lies in B, B contains A). The notation A rz. B (as the notation B ~ A) means that the set A is not a part of the set B (A is not a subset of the set B or A does not lie in B). The notation A ~ B does not exclude the case when the sets A and B may coincide, but the fact of the coincidence (= equality) of the sets itself is denoted by the formula A = B. Respectively, the fact of the non-coincidence of the sets is denoted by the formula A =I- B. Note that {x: x E A, P} denotes the set of all elements of a set A satisfying conditions, which are stated in P. Let us highlight that the notations sEA and s ~ A describe different situations and the notations sEA and {s} ~ A are equivalent. A set which contains no elements is called empty. Usually it is denoted by 0. The subsets 0 and A of a set A are called improper. All other subsets are called proper. We say that a mapping (a map, a function) f of a set A into a set B is defined, if for every element s of the set A there is associated an (unique) element t of the set B, which is denoted in this situation by f(s) (so t = f(s)) and which is called the value of the function (mapping) f at s or the image of s under the mapping f. The notation f : A ----+ B means that the function f is defined on the set A and takes values in B. The set A is called the domain of definition or the domain of f. Elements of the set A are called arguments of f. For M ~ A the symbol f(M) denotes the set of (all) values of the function f, which are taken at elements of the set M. For M ~ B the symbol f-1(M) denotes the sets of (all) elements of the set A for which the associated values of the function f belong to M. The set f(M) is called the image, and the set f -1 (M) is called the (full) preimage of the set M (under the mapping f). Often one uses mappings in descriptions of sets: a map f : A ----+ B distinguishes the set f(A) ~ B. Sometimes in such situations it is convenient not to forget how we have distinguished the set f(A), and after denoting the image of a E A by ta to use as a description of the mapping f the notation {t a : a E A}, which seems to list completely the set A and to point in this connection the image of each its element. Although the situation here may be completely reduced to those just described, for brevity we introd uce a new name: we call {ta: a E A} a family (of elements of the set B), which has the set A as an index set (each parti.cular element of the set
Topological and metric spaces.
3
A is called in this connection an index). We do not assume that elements of B which correspond to different indices are different too. In general the word 'family' is used in cases when we study the set J(A) and the mapping J plays an auxiliary role. Values of the function J are called elements of the family {ta: 0: E A}. In some sense the words 'set' and 'family' are equivalent. We mean that (with the above notation) the family mentioned describes the set J(A) and every set may be considered as a family (we can introduce here the structure mentioned by assuming that each element of the set is its own index). This allows us to transfer to sets notions and constructions introduced first for families, and vice versa. If Al S;; A, the family {t a : 0: E AI} is called a subfamily of the family
{to.:
0:
E
A}.
Values of the function J may be sets. In this case we speak about a family of sets, = {Xa: 0: E A} (we assume that A #- 0). The set of all x such that x E X a , at least for some 0: E A, is called the union of the family,. The set of all x such that x E Xa for every 0: E A is called the intersection of the family,. The union may be denoted by one of the following symbols: U" U{Xa: 0: E A}, UaEAX a . The intersection may be denoted by n" n{Xa: 0: E A}, naEAX a . A finite family, as a finite set, may be described by a complete enumeration of its elements. For instance, when the index set is the initial segment of the set of nat ural numbers A = {1, ... , n}, a family {Xa: 0: E A} may be denoted by {Xl,'" ,Xn} too. Its union and intersection may be denoted by U{XI"",Xn} = U?=IX i = Xl U",UXn and n{XI , ... , Xn} = n~IXi = Xl n··· n X n, respectively. If the index set A is the set of natural numbers N = {1, 2, ... } a family {Xa : 0: E A} may be denoted by {Xi: i = 1,2, ... } too. Its union and intersection may be denoted by U~l Xi and n~l Xi, respectively. Let A and B be arbitrary sets. The difference of the sets A and B (or the complement of the set B in A) is the set A \ B of (all) elements of the set A which do not belong to B. The product (or the Cartesian product) of a family of sets (of 'factors ') , = {Xa: 0: E A} is the set of all functions f : A --+ U" which satisfy the condition: J(o:) E Xa for each 0: E A. Such a product may be denoted by IT" or IT{Xa: 0: E A}, or ITaEA Xa' The product of a finite family {XI"",Xn} may also be denoted by IT{XI"",Xn}' or IT~=IXkl or Xl x '" X X n . The product of a family {Xk: k = 1,2, ... } may be denoted by IT;;:1 X k . In order to understand the contents of the notion introduced let us consider the particular case when the set of indices A consists of two elements: A = {I, 2}. An arbitrary element of the product Xl x X 2 associates to the first index an element Xl of the set Xl' It associates to the second index an
4
CHAPTER 1
element X2 of the set X 2 . So it may be described completely by the ordered pair {Xl, X2}' Vice versa: each such pair determines uniquely an element of the product Xl x X 2 . Thus when we introduce coordinates in the plane we identify the plane with the product of two real lines. This description may be expanded in a natural way to products of arbitrary finite number of factors. Let X and A be sets. Let Xu = X for every a E A. Usually we denote the product IT {Xu: a E A} by X A and we call it a power of the set X. Often in the notation X A the symbol of the set A is replaced by a symbol denoting its cardinality (see below). Following this understanding ]R2 denotes the Euclidean plane and ]Rn denotes the Euclidean space of dimension n. If any (for instance, algebraic) operation * is defined on a set X, then its action may be extended to subsets of X: for A, B ~ X we put A * B = {a * b: a E A, b E B}. Thus we define the sum A + B of subsets A, B of the real line or a vector space, a product AB of subsets A, B of the line or of the space of matrices, etc .. Similarly, we extend the action of operations which link elements of different sets. For instance, following this method the notion of the product of a matrix and a column is extended to sets of matrices and sets of columns, etc .. Notice the following simple propositions. Theorem 1.1 (de Morgan formulae). Let, be a family of subsets of a setX. ThenX\U,=n{X\H: HE,},X\n,=U{X\H: HE,} . • Theorem 1.2. (U{Xu: a E A})n(U{Y,6: (3 E B}) = U{XunY,6: {a,{3} E A x B}, (n{Xu: a E A}) U (n{Y,6: (3 E B}) = n{Xu U Y,6 : {a,{3} E A x B}. • Theorem 1.3. Let {Xu: a E A} be a family of subsets of a set X. Let J be a mapping of the set X into the set Y. Then J(U{Xu: a E A}) = U{J(Xu): a E A}.
•
Notice that in the last theorem we cannot change the symbols of the union to the symbols of the intersection: for instance, often we meet a situation where two sets do not intersect but the intersection of its images is nonempty (=f: 0). Theorem 1.4. Let {Y,6: {3 E B} be a family of subsets of a set Y. Let f be a mapping of a set X into the set Y. Then f-l(u{Y,6: (3 E B}) and
=
U{J-I(y,6): {3 E B}
•
Assume that A ~ X and f : A ----+ Y. Let -rr(f) = A (the domain of definition (or domain) of the mapping I), Gr(f) = {{x, f(x)}: x E A} ~
Topological and metric spaces.
5
x X Y (the graph of the mapping f) and Im(f) = f(A) (the set of values of the mapping f). A mapping f : A -+ B is called: injective, if images of two different elements of A are different (i.e., f(8) = f(t) implies 8 = t), surjective, if f(A) = B (i.e., every element of the set B is a value of the mapping f for some value of argument tEA), bijective or one to one, if it is injective and surjective at once (i.e., every element of the set B is associated to one and only one element of the set A), constant, if f(8) = f(t) for every 8, tEA, identity, if f(8) = 8 for every 8 E A (they write f(8) == 8 too). Often instead of 'f is a surjective mapping of a set A into a set B' they say: 'f is a mapping of a set A onto a set B'. For a bijective mapping f : A -+ B we can define the inverse mapping f- 1 : B 8
-+
A :
= f-l(t) if and only if t = f(s).
Sometimes it is convenient to consider a mapping which associates to points of its domain of definition A not elements, but subsets of the set B in such a manner as when we keep the terminology and notation introduced for ordinary mappings. Such a correspondence is called a multi-valued mapping. At the level of this definition in practice there are no differences between multi-valued mappings and the ordinary ones, which were discussed before (in order to distinguish them from multi-valued mappings we shall call them single valued): Every multi-valued mapping of a set A into a set B may be considered as a single valued mapping into the set of all subsets of the set B. Differences appear when we consider any supplementary structures on the sets A and B; for instance, if they are topological or vector spaces. A description of a structure on the set of all subsets of B, which corresponds strictly to the question under consideration, and its use in solving of problems may turn out to be too clumsy, and the use of the language of multi-valued mappings may be here more reasonable. So we shall say that a multi-valued mapping F of a set A into a set B is defined, if for every element sEA a subset F( s) of the set B is associated. Notice that unless otherwise stipulated we do not assume that values of the mapping are nonempty. Just as in the case of a single valued mapping the notation F : A -+ B denotes that F is a (multi-valued) mapping of the set A into the set B. For M ~ A the set F(M) = U{F(s): s E M} is called the image of the set M (under the mapping F). The symbol FIM denotes the restriction of the mapping F to the set M, i.e., it denotes the mapping G : M -+ B defined by G(s) = F(s). For M <;;; B the set F- 1 (M) =
CHAPTER 1
6
{s: sEA, F( s) n M i- 0} is called the preimage of the set M (under the mapping F). If there are two (multi-valued) mappings F : A ---) B and G : B ---) C, then we can define their composition GF : A ---) C by the formula GF(s) = G(F(s)). It is natural to extend this operation to arbitrary finite number of mappings if only values of each of them lie in the domain of definition of the next one. Every single valued mapping f : A ---) B may be considered as a multivalued one, which associates to a point sEA the singleton {f(s)}. Following this rule every multi-valued mapping which has singletons as values is generated by a single valued one. In our account we will not meet situations when a single valued mapping and the generated multi-valued one describe relations with different meanings. It will be convenient for us to identify them. Under this identifying the notions of image and preimage of a set for a single valued mapping which were introduced before correspond to the analogous notions which are defined for the generated multi-valued mapping. The introduced notions of composition and restriction may be applied to single valued mappings, moreover a composition of single valued mappings is a single valued mapping, a composition of injective mappings is an injective mapping, a composition of surjective mappings is a surjective mapping, a composition of bijective mappings is a bijective mapping, a restriction of a single valued mapping is a single valued mapping, etc .. When we speak about finite sets we know how to answer the question which of the sets under consideration contains more elements. In order to do that, it is enough to count the number of elements in each of them and then to compare results. When we speak about infinite sets such a procedure become impossible (at least until we define analogs of natural numbers and develop a corresponding theory). A full analysis of a situation arising may lead a very long way from the main topic of this book, but it is helpful to be oriented in some related questions. So we will introduce and discuss some related notions. We say that sets A and B are of equal cardinality (either equivalent, or the cardinalities of the sets A and Bare equal) and we write IAI = IBI, if there exists a bijective mapping of the set A onto the set B. We say that the cardinality of the set A does not exceed the cardinality of the set B and we write IAI ~ IBI if there exists an injective mapping of the set A into the set B. Theorem 1.5. Let A i- 0. Then IAI ~ IBI if and only if there exist a surjective mapping 9 : B ---) A. Proof. 1. If IAI ~ IBI then there exist an injective mapping f : A ---) B. Fix an arbitrary point a E A. Let g(t) = {
f-1(t)
a
it t E f(A), if t E B \ f(A).
Topological and metric spaces.
7
This formula defines the surjective mapping 9 : B ---t A. II. Let a surjective mapping 9 : B ---t A exist. For every sEA the set 9- 1 (s) is nonempty. Fix an arbitrary point f(s) of this set. So we have defined the injective mapping f : A ---t B. The theorem is proved. • Theorem 1.6. Let IAI ~ IBI and IBI ~ IAI· Then IAI = IBI· See [AI]. • A class of all sets equivalent to a fixed one is called a cardinality. In the case of infinite sets the notion of cardinality extends the notion of number of elements. Amongst all infinite cardinalities there is one minimal. This is the cardinality of the set of positive integers N = {I, 2, 3, ... } (see [AI]). Sets which are equivalent to the set of positive integers are called countable. A set is called at most countable if its cardinality does not exceed the cardinality of the set of positive integers. By virtue of the previous remark such sets are finite or countable. However, by misuse of language following a fully formed tradition, in some cases 'countable' is said instead of 'at most countable'. So below we shall say 'countable base' of a topological space assuming that the base in question may be finite. Families which have the set N of positive integers, or any subset of it, as an index set are called sequences. Their subfamilies are called subsequences. As a rule we will deal with sequences with infinite index sets. When we use the word 'sequence' without the adjective 'finite' we mean a sequence with infinite index sets. Let A be a infinite subset of the set N. For k E A let
Corollary. A product of a finite number of at most countable sets is at "nost countable. •
CHAPTER 1
8
Notice that if we have a product of a countable number of two-element sets its cardinality turns out to be uncountable (i.e., its cardinality is greater than the cardinality of the set N), see [AI]. Example 1.1. The set Z = {a, ±1, ±2, ... } of all integers is countable: we apply Theorem 1.7 to the representation
Z = {a} U {I, 2, ... } U {-I, -2, ... }. Example 1.2. The set Q of all rational numbers is countable: by Corollary of Theorem 1.7 the set M = Z x (Z \ {o}) is countable; the mapping rp : M ----t Q, rp(p, q) = p / q, is surjective and we obtain what was required from Theorem 1.5. 2. Topological spaces Let us start with the definition of the basic notion. Let X be an arbitrary set. A set T of subsets of X is called a topology on the set X, if 1) the sets X and 0 belong to T, 2) the union of every family of elements of T belongs to T, 3) the intersection of every finite family of elements of T belongs to T. Sometimes it is convenient not to use the above list of conditions but an equivalent list consisting of conditions 1), 2) and the following condition 3') the intersection of every pair of elements of the set T belongs to T. The equivalence of this two approach to the definition of a topology is obvious (condition 3') is a particular case of the condition 3) and the deduction of 3) from 1) and 3') may be realized by an evident induction). A couple (X, T), where X is a set and T is a topology on the set X, is called a topological space. Following a fully formed tradition, elements of the underlying set X are called points of the topological space (X, T). More frequently the space itself is denoted by the same symbol as the underlying set, i.e., instead of (X, T) we write X. Of course, this is possible only in cases when we consider only a unique topology on X. Usually, elements of the topology T are called open sets (of the topological space under consideration). In this language the conditions 1), 2), 3), 3') may be stated as follows: 1) the set X and the empty set are open, 2) the union of every family of open sets is open, 3) the intersection of every finite family of open sets is open, 3') the intersection of every two open sets is open. We can define many topologies on a given set X, except the trivial cases of the empty set and of an one point set (singleton) X. Notice that the number of topologies on a finite set is finite and the larger the set is the larger the number of topologies is. Naturally, the question arises of
Topological and metric spaces.
9
how to describe a particular topology. How we can select in the full set of topologies on a set X an unique topology which corresponds to a problem under consideration? Notice first that every set X underlies two topologies admitting trivial descriptions. There are: the discrete topology Td, the set of all subsets of the set X; the anti-discrete topology Ta = {X,0}. In both cases the fulfilment of the above axioms is obvious. Evidently every third topology T (on the same set X) lies between the two mentioned: Ta ~ T ~ Td·
Before discussing a rather extensive approach to the introduction of a topology on a set let us introduce the following notion. Let M be a (nonempty) set. Let a be a set of subsets of M. The set a is called directed if it is nonempty, does not contain the empty set (as its element), and for every pair of elements Al and A2 there exists an element A E a lying in the intersection of Al and A 2: A ~ Al n A 2. Let us mention some examples of directed sets. Let lR denote the real line. Example 2.1. For x E lR put Ax = [x, (0). Evidently for x,y E lR we have Ax nAy = Amax{x,y}. The family a = {Ax: x E IR} (is nonempty and) does not contain the empty set. Thus the set a is directed. Bearing in mind the geometry of the situation we san say that the set is directed to 00: under the decreasing of the set Ax, when x varies, the set 'shrinks' to 00.
Example 2.2. For x > 0 put Ax = [-x,x]. Evidently the set {Ax: x > O} is directed. Bearing in mind the geometry of the situation we san say that the set is directed to 0: under the decreasing of the set Ax, when x varies, the set 'shrinks' to o. Returning to the definition of a directed set notice that the relation of the inclusion Al ~ A2 defines an ordering on the set a (see below §8). So the condition in the definition of a directed set may be expressed as follows: every two elements of a are succeeded by a third one. As a simplest example of a directed set we can take a decreasing sequence of sets: Al "2 A2 "2 A3 "2 .... However, it does not always happen in applications which we have in mind that the elements of the set a are ranged in a decreasing chain. A directed set is called a filter if it satisfies a supplementary condition: if A E a and A ~ B ~ X, then B E a. Every directed set may be completed to a filter. However, this completion is not unique. Amongst all completions of a directed set a to a filter there is a minimal one. Denote it by F(a). A set B ~ X belongs to the filter F(a) if and only if there exists an element A of the set a for which A ~ B. We say that the filter F(a) is generated by
10
CHAPTER 1
the directed set CY. Properties of two directed sets turn out to be equivalent in many respects if their generated filters coincide. We say that a set r of subsets of a set X is directed to a point x E X if it is directed and each element of the set r contains the point x. In this case the belonging to an element of r may be considered as a measure of proximity to the point x: the smaller the element of the family r we take the nearer to the point x we are situated. Use this comparison to select subsets G of the set X, such that for every point x of G the set G contains all points which are quite near to the point x. To each point x of a set X let there be associated a set r x directed to the point x. Call a subset G of the set X 'open' (with respect to the family {r x: x E X}), if for each point s of the set G there exists an element A of the set rs lying in the set G: x E A ~ G. Theorem 2.1. The set T of (all) 'open' sets is a topology on the set X. Proof. The proof consists in checking the fulfilment of the above axioms. 1. 0 E T. This is obvious. X E T. For each point x E X the set r x is nonempty. Thus it contains some element A. Since A ~ X, this fact means that the set X is 'open'. 2. Let , ~ T and x E G = U,. There exists an element G 1 of the family, containing the point x. Since the set G 1 is 'open', there exists an element A of the set rx lying in G I . Thus x E A ~ G 1 ~ G. This means that the set G is 'open'. 3'. Let G 1 , G 2 E T and x E G = G I n G 2 . By virtue of the 'openness' of the sets G 1 and G 2 there are sets AI, A2 E r x such that Al ~ G I and A2 ~ G 2. The directedness of the set r x implies the existence of a set A E r x lying in the intersection Al n A 2 . Thus x E A ~ Al n A2 ~ G I n G 2 = G. This means the 'openness' of the set G. The theorem is proved. • It is more convenient to discuss properties of the topology T so described when the sets r x , x E X, themselves consist of 'open' sets; i.e., for every point x EX, for every element A of the set r x and for every point y of the set A there exists an element B of the set r y such that B ~ A. In this case the family {r x: x E X} is called a system of neighborhoods (on the set X). The mentioned description of the topology T is called definition of a topology by pointing of a system of neighborhoods. Example 2.3. Let X be an arbitrary set. For x E X let r x = {{x}}.
Evidently the mentioned system of neighborhoods generates the discrete topology on the set X. Example 2.4. Let r x = {[x, x + c): c > O} for x E R The topological space (with real line lR as its carrier set) defined by this system of neighborhoods is called an arrow or ZorgenJrei line (Figure 1.1).
Topological and metric spaces.
X
11
X+8
Figure 1.1
Example 2.5. Let X = {(s, t): s, t E JR., t ~ O} be the upper half plane including its bounding line. For a point x = (s, t) lying in the open half plane Xo = {(s, t): s, t E JR., t> O}, denote by fx the set of all open (i.e., without bounding circle) discs which lie in Xo and which have the point x as its center. For a point x = (s, t) lying in the bounding line t = 0 denote by f x the set of all sets of the form of {x} U D, where an open disc D is contained in the open half plane Xo and its bounding circle is tangent to the line t = 0 at the point x. This topological space is called the 'Nemytski plane' (Figure 1.2).
"-
/
\
/
I
\ I
I \ ......
-- -
/
/
/
/'
-- -
"-
\
I
\
I \ ......
/ ..........
/'
Figure 1.2
The description of a topology by pointing a system of neighborhoods is suitable not only in the introduction of a new topology but also in the investigation of one already introduced. In this latter case the following notion is convenient. A family f3 of open subsets of a topological space X is called a base at a point x E X if all elements of the family f3 contain the point x, and for every open set G :7 x of the space X there exists an element of the family f3 lying in G. Let us propose to the reader as a fruitful exercise proving that if we associate to every point x of a topological space (X, T) a base f3x at that point, then the family {f3x: x E X} is a system of neighborhoods and the topology generated by f3 coincides with the initial topology T. A family f3 of open subsets of a topological space X is called a base (of the space X, of the topology T) if for each point x of the space X and for
12
CHAPTER 1
each open set G 3 x there exists an element B of the family (3 such that x E B ~ G. Every base (3 (of a space X) satisfies the following two conditions: A. u(3 = X (because the set X is open); B. For every two elements Bl and B2 of the base (3 and for every point x E Bl n B2 there exists an element B of the base (3 such that x E B ~ Bl n B2 (because the set Bl n B2 is open). It is naturally to raise the question of whether every family (3 satisfying conditions A and B is a base of a topology on the set X and whether such a topology is unique. In order to give an affirmative answer to the first part of the question let us notice that under such assumptions the families (3x = {B: x E B E (3} constitute a system of neighborhoods on the set and the family (3 is a base of the topology generated by the family. This follows immediately from the definition of an 'open' set. An affirmative answer to the second part of the question follows from the definition of a base: A set is open if and only if it may be represented as the union of a family of elements of the base. Let us introduce one more notion that is close to those considered. A family (3 of open subsets of a topological space (X, r) is called a subbase (of the space X, of the topology r, respectively, sub-base at a point x of the space X), if the intersections of all possible finite subfamilies of the family (3 constitute a base of the space X (respectively, a base at the point x).
3. Some notions related to a topology Let x be a point and M be a subset of a topological space X. A set U is called a neighborhood of the point x (respectively, of the set M) if it is open and it contains the point x (respectively, the set M). Sometimes the word 'neighborhood' is given a wider meaning and they call a set H a neighborhood of the point x (respectively, of the set M) if there exists an open set U ~ H containing the point x (respectively, the set M). We will not use such terminology. An open set is a neighborhood of each of its points and of every subset. An intersection of a finite number of neighborhoods is also a neighborhood. Notice that the notion introduced is compatible with the use of the words 'system of neighborhoods' in the previous section. The point x is called an interior point of the set M (in the topological space X) if there exists a neighborhood of the point x lying in the set M. The set of all interior points of the set M is called the interior of the set M. It is denoted by (M). If it is necessary to distinguish the enveloping space the interior is denoted by (M)x. Evidently (M) ~ M, (Ml n M 2 ) =
Topological and metric spaces.
13
(Ml ) n (M2) and ((M)) = (M). Evidently the interior of every subset M of a topological space is an open set and it may be represented as the union of all open sets lying in M. A set is open if and only if it coincides with its interior. A point x is called an adherent point of a set M if every neighborhood of the point x intersects (or meets) the set M (i.e., the intersection of the neighborhood and the set M is nonempty). The set of all adherent points of a set M is called the closure of the set M. It is denoted by [M] or (if it is necessary to distinguish the enveloping space) by [M] x . Of course, every point of a set M is an adherent point of the set, i.e., M <;;;; [M]. But a set may have adherent points which do not belong to it. Evidently if Ml <;;;; M 2, then (M1 ) <;;;; (M2) and [Ml] <;;;; [M2]. Theorem 3.1. For every set M we have [[M]] = [M]. • Theorem 3.2. For every finite family of sets {Ml' ... ' Md we have
[U{Ml' ... ' Md]
=
u{[Md,···, [Mk ]}.
See [En,§1.1] or [AI]. • Evidently if f3 is a base at a point x E X and M <;;;; X, then: x E (M) if and only if some element of the base f3 lies in M; x E [M] if and only if every element of f3 intersects M. Evidently X \ [M] = (X \ M). A set M is called closed if [M] = M. Theorem 3.3. A set is closed if and only if the complement to it (in the space under consideration) is open. See [En,§1.1] or [AI]. • The following assertion is a direct consequence of Theorems 3.3 and 1.1 and properties of open sets. Theorem 3.4. Sets X and 0 are closed. The intersection of every family of closed sets is closed. The union of every finite family of closed sets is closed. The first assertion of the theorem is obvious (i.e., it follows immediately from definitions). The last one may be also obtained from Theorem 3.2. By Theorem 3.1 the closure of every set is a closed set. It is useful to have in mind that the closure of every set coincides with the intersection of all closed sets containing the set in question (see [En,§1.1] or [AI]). • The set [M] \ (M) is called the boundary of the set M. It is denoted by aM or (if it is necessary to distinguish the enveloping space) by ax M. The equality X \ [M] = (X \ M) implies that aM = [M]
n [X \
M] = X \ ((M) U (X \ M)).
Let a set Y lie in a topological space (X, T). It easy to prove that the family Ty = {u n Y: U E T} is a topology on the set Y. We say that the
14
CHAPTER 1
topology Ty is induced on the set Y by the topology T. A set Y ~ X with the topology Ty is called a subspace of the space (X, T). The induced topology is defined so that a set M ~ Y is open in the subspace Y if and only if there exists an open subset U of the space X such that M = U n Y. It easy to prove that a set M ~ Y is closed in the subspace Y if and only if there exists a closed subset F of the space X such that M = F n Y. For an arbitrary set M ~ Y we have [M]y = [M]x n Y. Notice important particular cases. If the set Y is open in the space X then the set M ~ Y is open in Y if and only if it is open in X. If the set Y is closed in X then the set M ~ Y is closed in Y if and only if it is closed in X. If f3 is a base at a point y E Y for the topology of the space X (respectively, f3 is a base of X), then the family {U n Y: U E f3} is a base at the point y in the space Y (respectively, a base of the space Y). If X ~ Y ~ Z then Tz = (Ty)z, i.e., the topology induced on the set Z by the topology of the space X coincides with the topology induced on Z by the topology of the subspace Y. 4. Metric spaces
Let X be an arbitrary set. A real function p : X x X -+ lR is called the metric or the distance on the set X if for every x, y, z E X: 1) p(x, y) = p(y, x) ~ 0; 2) p(x, y) = 0 if and only if x = y; 3) p(x, z) :::; p(x, y) + p(y, z). A couple (X,p), where X is a set and p is a metric on the set X, is called a metric space. As in the case of a topological space, elements of the set X are called points of the space (X, p). If it does not lead to a misunderstanding the space itself is denoted by same symbol as its carrier set X. The plane with the usual distance on it is an example of a metric space. The condition 3) corresponds to the following well known property of the plane:
IACI :::;
IABI
+ IBCI
for an arbitrary triangle ABC.
In this connection condition 3) is called usually the 'triangle inequality'. Example 4.1. Introduce on the real line lR the distance p(x, y) = Ix-yl. The fulfilment of all conditions is obvious. Example 4.2. The 'Euclidean' distance on the space lRn is defined by the formula
15
Topological and metric spaces.
where x = (Xl'···' Xn), Y = (YI, ... , Yn) E JRn. The verification of the fulfilment of the triangle inequality for the function p may be found in [AI]. For n = 1 we obtain the metric of Example 4.l. Example 4.3. Let X be an arbitrary set. For x, Y E X put
p(X,y) =
{~
if X i= y, if X = y.
The fulfilment of all axioms of a metric for this function p is obvious. Example 4.4. Let X be an arbitrary set and (Y, p) be a metric space. Denote by B(X, Y) the set of all bounded mappings of the set X into the space Y, i.e., the set of all mappings J : X --+ Y, for which sup{p(j(s),J(t)): s,t E X} < 00. For J,g E B(X, Y) put d(j,g) = sup{p(j(s),g(s)): SEX}. The verification the fulfilment of the triangle inequality for the function d may be found in [AI]. Example 4.5. With the notation of the previous example let X = {I, ... , n} and Y = R Identify an element J of the space B(X, JR) with the sequence {J(I), .. . ,J(n)}. In this way we define a one to one correspondence between the sets B(X, JR) and JRn. The metric on the space JRn which we obtain as a result of this correspondence by the formula of Example 4.4 differs from the Euclidean metric (see Example 4.2). The new metric is defined by the equality
d(x,y)=max{lxk-Ykl: k=I, ... ,n}, where x = (Xl, ... ,x n ), Y = (YI, ... ,Yn)· The subsets
O(x, E) = {t: t E X, p(X, t) < C},
where
E> 0,
Of (x, E) = {t: t E X, p(x, t) :::; E},
where
E ~ 0,
and of a metric space (X, p) are called, respectively, open and closed balls of the center X and the radius E. The set O(x, E) is also called the E-neighborhood of the point x. Often it is denoted by O,x. Open and closed balls O(x, E) and Of (x, E) contain the point x, and therefore they are nonempty. For each point X E X the set rx = {O(X,E): E > O} is directed because O(X,E) nO(x,8) = O(x,min{E,8}). If we add to the fact above mentioned that the set r x is nonempty (it contains, for instance, the ball O(x, 1)) then we obtain its directed ness to ~he point x. The topology generated by the mentioned system {r x : X EX} IS called the topology the metric space (X, p). Thus a set G ~ X is open in
16
CHAPTER 1
the metric space (X, p) if for each point t of the set G the E-neighborhood of the point t lies in the set G for some E > O. Example 4.6. For the real line (see Example 4.1)
+ E),
O(X,E) = (x - E,X O,(x, E)
=
[x - E, X + E]
and every (open) interval is an open set. Example 4.1. For the plane with the Euclidean distance (see Example 4.2) a ball of center x and radius E is the disc with center x and radius E, excluding its boundary circle in the case of the open ball and including the circle in the case of the closed ball. Example 4.8. For the three-dimensional space with the Euclidean distance the notions of a ball with center x and radius E coincide in the meaning of our definition and in the common geometric meaning. The open ball does not include its boundary sphere and the closed ball does include it. Example 4.9. For the space (X, p) from Example 4.3 if if
E ~
1,
>
1.
E
if if
E
< 1,
E ~
1.
°
In particular, (x, 1) = {x}. This implies that the topology of this space is discrete (see also Example 2.3). Example 4.10. On the plane with the metric of Example 4.5 the ball ,/
/ I
I
...........
(S\{I\
x
\
'" '-.... ..........
I
= (x - E, X + E)
has the form of a square (Figure 1.3).
\
~I
_-
/ ,/
Figure 1.4
- s I < E, Iy - t I < c} (y - E, Y + E)
= {( s, t): s, t E~, X
Ix
\
'-
\
Figure 1.3
o ((x, y), E)
--- /-"
I
I I I I I I I I I
/
/
/
Topological and metric spaces.
17
In order not to create false perceptions it is helpful to have in mind that although for the Euclidean metric O,(x, c) = [O(x, c)] always, for the metric of Examples 4.3 and 4.9 we have O,(x, c) i= [O(x, c)] for c = l. Theorem 4.1. Every open ball is an open set. Proof. Let s E O(x,c). Since p(x,s) < c, the number 8 = c - p(x,y) is positive. The triangle inequality implies O(x, 8) ~ O(x, c) (Figure 1.4), that gives the requirement. The theorem is proved. • The number p(A, B) = inf{p(s, t): sEA, t E B} is called the distance between (nonempty) subsets A and B of a metric space (X, p). Evidently p( {s}, {t}) = p(s, t) for all points s, t E X. If SEX and B ~ X, then instead of p( {s}, B) we usually write p(s, B) or p(B, s). Remark 4.1. Evidently a point x of a metric space (X, p) belongs to the closure of a set M ~ X if and only if p(x, M) = O. Let Y be a subset of a metric space (X, p). The function py : Y X Y ~ [0,00), py(s, t) = p(s, t), is a metric on the set Y. We say that the metric py is induced on the set Y by the metric p. The metric space (Y, py) is called a subspace of the metric space (X, p). With this notation let y E Y and c > O. Denote by 0; y the c-neighborhood of the point y in the space (X, p). Denote by O'{ y the c-neighborhood of the point y in the space (Y, py). Evidently
(4.1) This immediately implies Theorem 4.2. Let a set M lie in a subspace Y of a metric space (X, p). The set M is open in the subspace (Y, py) if and only if there exists an open • subset U of the space (X, p) for which M = Y n U. In other words, the induced metric generates the induced topology.
5. Limits A family ~ = {XOt : a E A} is called a generalized sequence if some directed set ep of subsets of the index set A is specified. A sequence may be considered as a particular case of a generalized sequence. Naturally, in this case we take the set of positive integers N or an infinite subset of it as the index set A, and the set {A n { k, k + 1, k + 2, ... }: k = 1,2, ... } as ep. A point x E X is called a limit point of a set M ~ X in a topological space X if every neighborhood of the point x contains an infinite number of elements of the set M. A point x E X is called a limit point of a generalized sequence {XOt : Q E A} of elements of the space X (with respect to ep) if for every neighborhood Ox of the point x and for every set F E ep we have XOt E Ox for
18
CHAPTER 1
some a E F. In the case of a sequence the latter condition is equipotent with the condition: for each neighborhood Ox of the point x the set {k: k E A, Xk E Ox} infinite. A point x E X is called a limit of a generalized sequence {xc>: a E A} of elements of the space X (with respect to cp), iffor each neighborhood Ox of the point x there exists an element F of the set cp such that Xc> E Ox for every a E F. In the case of a sequence the last condition looks as follows: for each neighborhood Ox of the point x there exists ko such that Xk E Ox for every k E A, k ~ ko (in what follows this fact will be expressed briefly as: 'beginning with some k = ko '). The fact that the point x is a limit of a generalized sequence {xc>: a E A} may be also written Xc> ---+ x or x = limlf' xC>. Often in the last notation the symbol cp is replaced by other ones, see below. A topological space X is called Hausdorff (or a T2 space) if for every pair distinct points sand t of it there exist disjoint neighborhoods Os and Ot, respectively: Os n Ot = 0. Theorem 5.1. Every metric space is Hausdorff. Proof. Let x and Y be two different points of a metric space (X, p). Then OEX n OeY = 0 for E = tp(x, y). This follows immediately from the triangle inequality. The theorem is proved. • Theorem 5.2. A generalized sequence of points of a Hausdorff space cannot have more than one limit. Proof. Assume that a generalized sequence {xc>: a E A} of points of a Hausdorff space X has two different limits Yl and Y2 (with respect to a directed set cp). Let OYl and OY2 be disjoint neighborhoods of the points Yl and Y2' By the definition of a limit, for every i = 1,2 there exists a set Fi E cp such that Xc> E 0Yi for every a E F i . By virtue of the directedness of the set cp there exists an element F of cp lying in the intersection Fl n F 2 . For a E F the point Xc> needs to lie in OYI and in OY2 at once. But that is impossible because the sets OYI and OY2 are disjoint. So our assumption is false and the theorem is proved. • We noticed in §1 that after a corresponding re-numbering an arbitrary sequence turns into a sequence with the positive integers as the index set. Under this re-numbering the convergence of the sequence to a point (as well as the absence of this convergence) is kept. A subsequence of a convergent sequence converges to the same limit. An analogous assertion it may be given for a generalized sequence too. Theorem 5.3. Let a point x of a topological space X have (at most) a countable base. The point x belongs to the closure of a set M ~ X if and only if there exists a sequence a of points of the set M converging to the
Topological and metric spaces.
19
point x. The point x is a limit point of the set M if and only if the sequence
a may be composed by different points. Proof. 1. Let the point x E [M] possess a countable base f3 = {Ui : i = 1,2, ... }. For every i = 1,2, ... the intersection U 1 n ... n Ui n M is nonempty. For every i = 1,2, ... fix a point Xi E Ul n ... n Ui n M. We obtain a sequence of points of the set M, which converges to the point x. II. Let the point x be a limit of a sequence a of points of the set M. Then every neighborhood of the point x contains at least one element of the sequence a, and hence it contains a point of the set M. So x E [M]. If the sequence a of the previous reasoning consists of pairwise different points then by analogous arguments an arbitrary neighborhood of the point x contains an infinite number of points of the set M. Therefore the point x is a limit point of the set M. III. If x is a limit point of the set M then we realize the choice of the point Xi in I sequentially under the fulfilment of the supplementary condition Xi E (U l n··· n Ui n M) \ {Xl,"" Xi-d. The theorem is proved. • Every point X of a metric space possesses a countable base, for instance, {O(X,2-i): i = 1,2, ... }. Therefore Theorem 5.3 may be immediately applied to metric spaces. In fact we may describe all topological properties of metric spaces in the language of convergent sequences (although it is not always expedient). Introduce now two notions that are close to the notion just considered of a limit of a generalized sequence of points. Let now
= {X,,: a E A} be a generalized sequence of subsets of a topological space X. The upper topological limit of the generalized sequence with respect to the directed set cp is defined as the set of all points x of the space X satisfying the condition: for every neighborhood Ox of the point x in the space X and for every F such that X" n Ox =f. 0.
FE cp there exists an index a E
The lower topological limit of the generalized sequence with respect to cp is defined as the set of all points x of the space X satisfying the condition: for every neighborhood Ox of the point x in the space X there exists an element F E cp such that X" n Ox =f. 0 for every a E F. Upper and lower topological limits are denoted usually by lim top sup and lim top inf, respectively, with the addition of a symbol for the generalized sequence in question or its common term, at the right of this symbol and of the symbol for the set cp, or of an any substituting symbol at the
CHAPTER 1
20
bottom. When we speak about a sequence we usually write in the latter case k ----+ 00. Evidently lim top inf'l' {X" : a E A} ~ n{[u{X,,: a E F}]: FE cp} = lim top sUP'l' {X" : a E A}. Example 5.1. Enumerate arbitrarily, but without repetitions, the set Q of all rational points of the real line JR (see Example 1.2): Q = {rk : k = 1,2, ... }. Then lim top infk--->oo{rd = 0 and lim top SUPk---> 00 {rd = JR. Example 5.2. Take as the space X the Euclidean plane JR 2 . Denote by Ak the set
Figure 1.6
Figure 1.5
for k = 2,4,6, . .. and the set
for k = 1,3,5, .... Denote by Bk the set
{2-k} x [0,00) for k = 2,4,6, ... and the set
Topological and metric spaces. for k
= 1,3, ....
21
Then lim top sup Ak
= {O}
x [0,2- 1 ],
k---+oo
limtopinfAk
=
{O} x [0,2- 2 ] (Figure 1.5),
k---+oo
lim top sup Bk = {O} x [0,00), k---+oo
lim top inf Bk = 0 (Figure 1.6). k---+oo
Example 5.3. For a non-decreasing sequence of sets Ai C;;; A2 C;;; .•• lim tOPSUpAk = lim topinf Ak = [U{Ak: k = 1,2, ... }]. k---+oo
k---+oo
For a non-increasing sequence of sets Bl "2 B2 "2 B3 "2 ...
limtopsupBk = limtopinfBk = n{[BkJ: k = 1,2, ... }. k---+oo
k---+oo
In particular, for a stationary sequence C k
== C
lim top sup Ck = lim top inf Ck = [C]. k---+oo
k---+oo
The upper (respectively, lower) limit of a generalized sequence with respect to a directed set 'P and with respect to a filter F( 'P) (see §2) coincide. This implies that if directed sets 'Pi and 'P2 generate the same filter then the upper (respectively, lower) topological limits with respect to them are equal. Often the fulfilment of the condition F( 'Pd = F( 'P2) may be checked quite easely: Evidently it is necessary and sufficient that for every element Fl of the set 'Pi there exists an element of the set 'P2 lying in F 1 , and for every element F2 of the set 'P2 there exists an element of 'Pi lying in F2. A topological space may appear as an index set of a generalized sequence. In this case the set 'P of all neighborhoods of an arbitrary point a of this space and an arbitrary base f3 at this point generate the same filter. Therefore the upper (respectively, lower) topological limits with respect to 'P and f3 coincide. In the notation of the limits these directed sets are denoted by a ----+ a, where a denotes a common index of the generalized sequence: lim top sUPa---+a Xa and etc .. If a is an adherent point of a set M, then what is more the families {UnM: U E 'P} and {UnM: U E f3} are directed and the filters generated by them on the set M coincide. Therefore Upper (respectively, lower) limits with respect to this directed sets coincide too. In the notation of limits these directed sets are usually denoted as a ----+ a, a EM. A point a of a topological space is called non-isolated (respectively, isolated), if the set {a} is not (respectively, is) open. Evidently a point a is a non-isolated point of a topological space A if and only if it is an adherent
22
CHAPTER 1
point of the set M = A \ {a}. In this case the families {U\ {a} : U E cp} and {U\ {a} : U E ,8} are directed and their generated filters coincide. Therefore upper (respectively, lower) limits with respect to them coincide too. In the notation of limits these directed sets are also denoted (side by side with the notation of the previous paragraph) as Q' - . a, Q' =/:- a. Sometimes (when this is mentioned in addition and does not cause misunderstandings because of other reasons) it may be written as the shortened notation Q' - . a. In the case of a metric space the notion of the upper limit of a sequence is clarified well by: Lemma 5.1. Let {Xk: k E A} be a sequence of subsets of a metric space X and x EX. The point x belongs to the set lim top SUPk-> 00 X k if and only if: (5.1) we can choose points
Xk
E Xb where k runs over an infinite subset {Xk: k E Ad converges to the point
Al of the set A, so that the sequence x.
Proof. Show that the membership x E lim top SUPk-> 00 X k implies the fulfilment of condition (5.1). Take as ki an arbitrary index k E A, for which X k =/:- 0 (such an index exists by virtue of the nonemptiness of the upper limit of the sequence in question). Take as Xkl an arbitrary point of the set X k1 . Let indices kl , ... , ki - I and points Xk 1 E X k1 , . .. , Xki_l E X ki _ 1 be fixed. The set {k: k E A, k > ki - I , X k n O(x, +) =/:- 0} is nonempty. Take as k i an arbitrary element of the latter set and take as Xki an arbitrary point of the set X ki nO(x,+). Put Al = {k i : i = 1,2, ... }. Evidently condition (5.1) is fulfilled. Condition (5.1) obviously implies the membership x E lim topsupXk. The lemma is proved. k->oo • A generalized sequence q, is called convergent (with respect to the directed set cp) if lim top sUPcp q, = lim top infcp q,. In this case the common value of the upper and lower limits is called the topological limit of q, (with respect to cp) and is denoted by lim topcp q" with versions of this symbol being analogous to those for the upper and lower limits. Evidently in the case when we consider a generalized sequence of points {xc;:: Q' E A} its convergence with respect to the directed set cp to a point x is equipotent to the equality lim top infcp {xc;:: Q' E A} = {x}. The sequences of Example 5.3 converge. Theorem 5.4. Every sequence of subsets of a topological space with a countable base contains a convergent subsequence. Proof. Let ,8 = {Uk: k = 1,2, ... } be a countable base of a topological space X. Let /0 = {X k : k E Ao} be a sequence of subsets of the space X. We will choose subsequences of the sequence /0. Let i = 1,2, ... and that the sequence / i - l = {Xk: k E Ai-d be already gived. At least one
Topological and metric spaces.
23
of the sets Bi = {k: k E A i- 1 , k > minA i_ 1 , X k n Ui i- 0} or B; = {k : k E A i - 1 , k > min A i - 1 X k n Ui = 0} is infinite. Let us take this set as Ai (any if both are infinite) and let us put = {X k: k E Ai}' Let A* = {min Ai: i = 1,2, ... } and = {Xk :. k E A*}. Show that the sequence converges. Since the lower topological limit of a sequence lies in the upper topological limit, we need to show only that an arbitrary point t of the upper limit of the sequence belongs to its lower topological limit. Let Ot be an arbitrary neighborhood of the point t. An element Ui of the base f3 containing the point t lies in Ot. As the point t belongs to the upper limit of the sequence and the set Ui is a neighborhood of the point t, then the sequence contains elements with arbitrarily great indices which meet the set Ui . But this is possible only in the unique case where in the construction of'i we have taken Ai = B i . Then all elements of the sequence ,., beginning with the element of the index min Ai, meet the set Ui too. Hence they meet the set Ot. This implies that the point t belongs to the lower topological limit of the sequence The theorem is proved. • Notice that in Theorem 5.4 we do not state the nonemptiness of the limit of the subsequence constructed.
'i
,* ,*
,*
,* ,*
,*.
6. Compactness and completeness
A family of sets , is called a cover of a set M if M S;;; U,. In this case every subfamily of the family, satisfying the same condition M S;;; U,o is called a subcover (of the cover, of the set M). A cover of a topological space X consisting of open (respectively, of closed) subsets of the space X is called open (respectively, closed). A topological space is called compact if every open cover of it contains a finite subcover. A compact Hausdorff space is called a compactum. As we do with all topological notions, the terms introduced will also be used for the description of topological properties of subsets of topological spaces. In this case we have in mind the induced topology on the subsets. When we define a compact set we can mean that in the definition we speak about the cover of the set in question by open subsets of the including space. The equivalence the notions obtained results easily from the definition of the induced topology. This implies, in particular, that the union of a finite family of compact subsets is compact. Evidently every finite set is compact. A subset M of a topological space is called relatively compact if the set [Mj is compact. Lemma 6.1. Let x be a point and M ~ x be a compact subset of a Hausdorff space X. Then there are disjoint neighborhoods Ox of the point x and OM of the set M, i.e., Ox n OM = 0.
,0
24
CHAPTER 1
Proof. Since the space X is Hausdorff we have the possibility of associating to every point t E M a neighborhood at of the point t and a neighbor hood G (t) of the point x such that at n G (t) = 0. Since the set M is compact its open cover {at: t E M} has a finite subcover {Ot l , ... ,Otk }. Evidently the sets Ox = n{G(t l ), ... , G(tk)} and OM = U{Ot l , ... , Ot k } satisfy the stated conditions. The lemma is proved. • Lemma 6.1 implies immediately Theorem 6.1. Every compact subset of a Hausdorff space is closed . • Thus every finite subset of a Hausdorff (in particular, of a metric) space is closed. Theorem 6.2. Every closed subset of a compact space is compact. See [En, Theorem 3.1.2] or [AI]. • Comparing these two theorems we obtain that a subset of a compactum is closed if and only if it is compact. We say that a family of subsets of a set X has the finite intersection property if the intersection of every its finite subfamily is nonempty. Theorem 6.3. A topological space is compact if and only if every family of its closed subsets with the finite intersection property has the nonempty intersection. See [En, Theorem 3.1.1] or [AI]. The proof is based on the use of Theorem 1.1. • Notice a particular case of this assertion. Corollary. A non-increasing sequence of nonempty closed subsets of a compact space has nonempty intersection. • Compactness is closely related to the property (6.1) every infinite subset of the space in question has a limit point. Theorem 6.4. Every compact space satisfies condition (6.1). See [AI] or [En, Theorem 3.1.23]. • In the case of a metric space Theorem 6.4 allows a converse and the following assertion holds. Theorem 6.5. A metric space satisfies condition (6.1) if and only if it is compact. See [AI] or [En, §4.3]. • Example 6.1. Every segment of the real line is compact. This fact is well known. Often it is stated in the form of one of the conditions (6.1), (6.2) or (6.3) (see below). Example 6.2. A subset of a discrete space (see Examples 4.3 and 4.9) is compact if and only if it is finite. ExaIllple 6.3. Let X = {a} UA, where a (j. A. Let r x = {{x}} if x E A, and r a be family of complements in X to arbitrary finite subsets of the set
25
Topological and metric spaces.
A. Consider the topology on X defined by the system of neighborhoods {f x: x EX}. Evidently this space is compact. Example 6.4. The 'Cantor perfect set'. For a segment [a, b], where a < b, denote by S([a, b]) the set
[a,a
b - a] + -3-
[
b-
a]
U b - -3-,b
(we reject the middle third). Let a set M be represented as a union of a finite family {II, ... ,Id of pairwise disjoint segments (prove as an exercise that if such a representation exists then it is unique, see also §2.6 beloW). Put S(M) = U{S(II), ... , S(h)} and Sk(M) = S(Sk-I(M)) for k = 2,3, .... Finally, we define the Cantor perfect set K by putting K = n{Sk([O, 1]): k = 1,2, ... }. Evidently we obtain as a result of the operation S a closed subset of the initial set. Thus we define the Cantor perfect set as the intersection of a decreasing sequence of nonempty closed sets lying in the segment [0, 1]. By the Corollary of Theorem 6.3 the set K is nonempty (however, it is also obvious without this reference). By Theorem 3.4 the set K is closed in the segment [0,1]. Therefore Theorem 6.2 implies that it is compact. Condition (6.1) is equipotent with the condition (6.2) every sequence of elements of the space in question has a limit point. (see [AI]). In the case of a metric space these conditions are equipotent with the condition (6.3) every sequence of elements of the space in question contains a convergent subsequence. (see [AI]). A sequence {x k: k E A} of points of a metric space (X, p) is called a fundamental, or Cauchy, sequence if for of every E > there exists a number N such that if i,j E A and i,j ~ N, then P(Xi' Xj) < E. Evidently every convergent sequence is fundamental. Every fundamental sequence cannot have more than one limit point. If a fundamental sequence possesses a limit point then the sequence converges to this point (see [AI]). A metric space is called complete if every its fundamental sequence converges. Theorem 6.4 and just made remarks imply Theorem 6.6. Every compact metric space is complete. • Example 6.5. Euclidean space )Rn is complete (see Example 4.2 and
°
[AID·
CHAPTER 1
26
Example 6.6. The space of Examples 4.3 and 4.9 is complete. This is easy to check. Example 6.7. Let a space Y be complete. Then the space B(X, Y) is complete (see Example 4.4 and [En, Theorem 4.3.14] or [AI]). A set M <;;; X is called dense in an open subset U of a topological space X, if U <;;; [M] or (what is equipotent) U <;;; ([MJ). Two extreme situations related with this notion will be important for us. A set M is called everywhere dense in X (or in brief, and completely according to the previous definition, is dense in X) if [M] = X. A set M is called nowhere dense in X if ([MJ) = 0. The latter definition and Theorem 3.1 imply immediately that a set M is nowhere dense in X if and only if the set [M] is nowhere dense in X. Theorem 6.7 (Baire). A nonempty open subset of a complete metric space cannot be represented as the union of an (at most) countable family of nowhere dense subsets.
See [En, Theorem 3.9.3] or [AI]. • In particular, a complete space itself may not be represented as the union of a countable family of nowhere dense subsets. Corollary. If the interior of the union of a countable family of closed subsets of a complete metric space is nonempty then at least one of sets of this family has the nonempty interior. • A topological space is called separable if it has an (at most) countable
everywhere dense subset. Theorem 6.8. Every compact metric space is separable. See [En,§4.3] or [AI]. Theorem 6.9. A metric space is separable if and only if it has an (at
•
most) countable base.
See [En,§4.3] or [AI]. • Theorem 6.10. Let, be a family of open subsets of a topological space
,0
with countable base. Then there exists an (at most) countable subfamily of the family" for which U,o = U,. Proof. Let f3 be an (at most) countable base of a space X. Denote by f30 the set of all elements U of the base f3 satisfying the condition
U <;;; G
for some
G E ,.
For every U E f30 fix an element a(U) of the family, containing U. Since /3 is a base, u/3o = U,. This equality and the obvious inclusion u/3o <;;; U,o <;;; u" where '0 = {a(U): U E /30}, imply the equality U,o = U,. Since the base /3 is countable the family /30 is countable. Hence the family is countable too. The theorem is proved. •
,0
Topological and metric spaces.
27
Corollary. Let, be a family of nonempty pairwise disjoint open subsets of a topological space with countable base. Then the family, is (at most) countable. • 7. Continuous mappings A mapping f : X - t Y of a topological space X into a topological space Y is called continuous at a point x E X if for each neighborhood Of (x) of its image f(x) in the space Y there exists a neighborhood Ox of the point x in the space X such that f(Ox) ~ Of (x). This notion may be expanded to multi-valued mappings in different ways. Below, the notion of an upper semicontinuous mapping will be most useful for us. A multi-valued mapping F : X - t Y of a topological space X into a topological space Y is called upper semicontinuous at a point x E X if for each neighborhood OF(x) of the set F(x) in the space Y there exists a neighborhood Ox of the point x in the space X such that F(Ox) ~ OF(x). Evidently the continuity of a single valued mapping f : X - t Y is equipotent to the equality limtopinft->xU(t)} = f(x). In the case of a multi-valued mapping the analogous equality describes a slightly different situation. Example 7.1. For t E IR put if t I- 0 ift = O. The mapping F satisfies the condition lim top inft->o F(t) = F(O), but it is not upper semicontinuous at the point t = O. The upper semicontinuity of a multi-valued mapping corresponds to an other notion of convergence of a generalized sequence = {Xex : a E A} of subsets of a topological space X to a set H ~ X (with respect to a directed set
28
CHAPTER 1
A sequence {ak: k = 1,2 ... } of elements of the set A is called cofinal to a directed set
= {Xo:: a E A} of subsets of a topological space X converges with respect to
not converge to the set H in the sense of (7.1). This means that for some neighborhood OH of H in every set F k , k = 1,2, ... , there exists an element ak such that XO: k \ OH -# 0. The sequence {ak: k = 1,2, ... } is cofinal to the set
= {Xa: a E A} of subsets of a regular space X converge in the sense of (7.1) with respect to a directed set
~ H. Proof. Our aim will be achieved when we show that every point x E X \ H does not belong to the set lim top sup'!' .
Topological and metric spaces.
29
Take disjoint neighborhoods Ox and OH of the point x and of the set H. By (7.1) there exists an element F of the set l{J such that Xc>: ~ OH for of every index a E F. Therefore Xc>: n Ox = 0. So x 1. lim top sup'!' . The theorem is proved. • Theorem 7.2. Let = {Xc>:: a E A} be a generalized sequence of subsets of a compactum X. Then the sequence converges (with respect to l{J) in the sense of (7.1) to the set H = lim top sup'!' . Proof. Let OH be an arbitrary neighborhood of the set H in the compactum X. Then every point x of the set X \ OH does not belong to the set lim top sup'!' . Therefore there exist a neighborhood of it Ox and a set F(x) E l{J such that for every index a E F(x) the intersection Xc>: n Ox is empty. By Theorem 6.2 the closed set X \ OH is compact. Its cover {Ox: x E X \ OH} has a finite subcover {OXI, ... , Oxd. By virtue of the directed ness of the set l{J there exists an element F of it lying in the intersection of the sets F(XI), ... , F(xd. If a E F then Xc>: n OXi = 0 for every i = 1, ... , k. Therefore Xc>: n (X \ OH) ~ Xc>: n (U~=I OXi) = U~=I (Xc>: n OXi) = 0. So Xc>: ~ OH. The theorem is proved. • Thus in the case of a compact space the convergence of a generalized sequence with respect to l{J in the sense of (7.1) to a closed set H is equipotent to the inclusion lim top sup'!' ~ H. This means, in particular, that for a multi-valued mapping F : X ----t Y with closed values of a topological space X into a compactum Y the upper semicontinuity of the mapping F at a point x E X is equipotent with the inclusion lim tOpSUPt-+x F(t) ~ F(x). By analogy with a base and a sub-base at a point a family f3 of open subsets of a topological space X is called a base (respectively, a subbase) of neighborhoods of a set H ~ X if every element of f3 contains the set H and for each neighborhood OH of the set H in the space X there exists an element U of the family f3 such that U ~ OH (respectively, there are elements UI , . . . ,Uk of the family f3 such that UI n ... n Uk ~ OH). Lemma 7.3. Let = {Xc>:: a E A} be a generalized sequence of subsets of a topological space X. Let f3 be a sub-base of neighborhoods of a set H ~ X. The sequence converges to the set H with respect to a directed set l{J in the sense of (7.1) if and only if for every element U of the sub-base f3 there exists a set F E l{J such that Xc>: ~ U for every index aEF. Proof. Necessity follows immediately from definitions. Sufficiency. Take an arbitrary neighborhood OH of the set H in the space X. By the definition of a sub-base there are sets UI , ... ,Uk E f3 such that UI n· .. n Uk ~ 0 H. For every i = 1, ... , k fix an element Fi of the set l{J such that Xc>: ~ Ui for every index a E F i . By virtue of the directedness of the set l{J there exists element F of it lying in the intersection FI n ... n F k •
30
CHAPTER 1
For every index a E F we have: XCI. <;;;; UI n··· n Uk <;;;; OH. The lemma is proved. • This lemma implies: Theorem 7.3. Let F : X -+ Y be a multi-valued mapping of a topological space X into a topological space Y, f31 be a base at a point x EX, f32 be a sub-base of neighborhoods of the set F(x) in Y. The mapping F is upper semicontinuous at x if and only if for every V E f32 there exists U E f31 such that F(U) <;;;; V. • In the case of a single valued mapping the notion of the continuity at a point coincides with the notion of the upper semicontinuity at the point. Therefore all assertions about upper semicontinuous (at a point) mappings may be immediately applied to continuous single valued mappings. In the case of a metric space £-neighborhoods of a point constitute its base. Therefore from Theorem 7.3 we obtain Corollary. A mapping f : X -+ Y of a metric space (X, PI) into a metric space (Y, P2) is continuous at a point x E X if and only if for every £ > 0 there exists 8 > 0 such that if t E X and PI (x, t) < 8, then P2 (f (X ), f (t )) < £. • In this way we usually state the definition of the continuity at a point of a real function of a real argument (see in Example 4.1 the metric in the real line). A mapping of a topological space X into a topological space Y is called continuous if it is continuous at every point of the space X. A multi-valued mapping of a topological space X into a topological space Y is called upper semicontinuous if it is upper semicontinuous at every point of the space X. We say that a mapping f of a metric space (X, PI) into a metric space (Y, P2) satisfies Lipschitz condition with Lipschitz constant a ~ 0 if P2(f(s),f(t)):::;; apI(s,t) for every pair of points s,t of the space X. Every such mapping is continuous. This follows easily from the Corollary of Theorem 7.3. We put there 8 = £/(a + 1) for arbitrary £ > O. Example 7.2. Let d be a metric on the product of two real lines, which is mentioned in Example 4.5. Let the mapping
l
+ tl·ls -
tl :::;; 2als - tl·
Topological and metric spaces.
31
In particular, all functions in Examples 7.2-4 are continuous. A mapping of a metric space into a metric space is called contractive if it satisfies the Lipschitz condition with a constant being less than 1. A point x of a set X is called a fixed point of a mapping f : X ---t X if f(x) = x.
Theorem 7.4. Every contractive mapping of a complete metric into itself space possesses a fixed point and such a point is unique. Proof. Let a mapping f of a complete metric space (X, p) into itself satisfy the Lipschitz condition with a constant a < 1. Take an arbitrary point Xl of the space X. For k = 2, 3, . .. define sequentially the points Xk by the formula Xk = f(Xk-d· Because 0 < a < 1, a k - l ---t 0 as k ---t 00. Since P(Xb Xk+i)
:s; :s;
+ P(Xk+l, Xk+2) + ... + P(Xk+i-l, Xk+i) P(Xk, Xk+l) + ap(xk' xk+d + ... + a i - l P(Xk, xk+d P(Xk, xk+d
1 - ai
= --P(Xk, Xk+l)
:s;
I-a a --P(Xk-I,Xk) I-a
:s;
1
--P(Xk' xk+d I-a
:s; ... :s;
a k-
l
--P(XI,X2) I-a
for every k = 1,2, ... and i = 1,2, ... , the sequence {Xk: k = 1,2, ... } is fundamental. Since the space X is complete this sequence converges to a point X E X. The sequence {f(Xk): k = 1,2, ... } is obtained from the sequence {Xk: k = 1,2, ... } by rejection of the first member. Therefore the limits of these sequences coincide. By Corollary of Lemma 7.1 or by Lemma 7.2 f(x) = limk--+oo f(Xk) = limk--+oo Xk = x, i.e., x is a fixed point of the mapping f. We have proved the existence of a fixed point. Prove its uniqueness. Let t be a fixed point of the mapping f too. As p(x, t) = p(f(x), f(t)) :s; ap(x, t), where 0 :s; a < 1, then p(x, t) = 0 and x = t. The theorem is proved. • Theorem 7.5. Let F : X ---t Y be a multi-valued mapping of a topological space X into a topological space Y. Then the following conditions are equipotent: a) the mapping F is upper semicontinuous, b) for every open set V of the space Y the set {x: x E X, F(x) S;;; V} is open in X, c) for every closed set H of the space Y the set F-I(H) is closed in X. Proof. a{:::}b. This follows immediately from the definitions. b{:::}c. This follows from Theorem 3.3. The theorem is proved. • Theorem 7.6. Let a multi-valued mapping F : X ---t Y of a topological space X into a topological space Y be upper semicontinuous on every element of a finite closed cover of the space X. Then the mapping F is upper semicontinuous (on the entire space X).
32
CHAPTER 1
Proof. Let I be the finite closed cover of the space X. The words 'F is upper semicontinuous on H(E I)' mean that the mapping F IH is upper semicontinuous. Our assertion follows from Theorem 7.5c, Theorem 3.4, the obvious equality
F-1(M) =
U{(FIH)
-1
(M): H
EI} ,
and properties of the induced topology. The theorem is proved. • Theorem 7.7. Let a multi-valued mapping F : X -> Y of a topological space X into a topological space Y be upper semicontinuous on every element of an open cover of the space X. Then the mapping F is upper semicontinuous (on the entire space X). Proof. The proof is analogous to the proof of the previous theorem .• With reference to Theorems 7.6 and 7.7 it is necessary to notice that a restriction of an upper semicontinuous mapping to a subspace is upper semicontinuous with respect to the induced topology. In addition a composition of upper semicontinuous mappings is upper semicontinuous. This may be easily deduced right from definitions or with the help of Theorem 7.5. Theorem 7.8. Let F : X -> Y be an upper semicontinuous multi-valued mapping of a compact space X into a topological space Y. Let for each point x E X the set F(x) be compact. Then the set F(X) is compact. Proof. Let I be an arbitrary cover of the set F(X) by open subsets of the space Y. For every x E X fix a finite subfamily Ix of the cover I, which covers the set F(x). By Theorem 7.5b the set Ox = {t: t E X, F(t) ~ U is open in X. Evidently the set contains the point x. Thus we have an open cover {Ox: x E X} of the compact space X. Select its finite sub cover {OXl,'''' Oxd. As F(OXi) ~ U for every i = 1, ... , k, then F(X) = U~=lF(OXi) ~ U~=l (U'xJ. Hence the finite subfamily U~=llXi of I covers the set F(X). This gives what was required. • Naturally this result may be applied to single valued mappings too: A continuous image of a compact space is a compact space. We can use the latter observation to give an example of an upper semicontinuous multivalued mapping which is remote from continuous single valued mappings. Corollary. Let f : X -> Y be a continuous (single valued) mapping of a compact space X into a Hausdorff space Y. Then the (multi-valued) mapping F : Y -> X, F(y) = f-1(y), is upper semicontinuous. In order to obtain the corollary we need use Theorem 7.5c. Let H be an arbitrary closed subset of X. By Theorem 6.2 the set H is compact. By Theorem 7.8 the set F-1(H) = f(H) is compact too. We obtain from Theorem 6.1 the closedness of the set F-1(H), what was required. • Finish this section by a small remark. We will have an opportunity to use it below.
,X}
,X;
Topological and metric spaces.
33
Lemma 7.4. Let J and g be continuous (single valued) mappings oj a topological space X into a Hausdorff space Y. Then the set (oj points of coincidence of the mappings f and g) M = {x: x E X, J(x) = g(x)} is closed in X. Proof. Take an arbitrary point t E X \ M. Since the points f(t) and g(t) are different and the space Y is Hausdorff, there exist disjoint neighborhoods U and V of the points J(t) and get), respectively. By virtue of the continuity of the mappings f and g there exists a neighborhood at of the point t such that f(Ot) ~ U and g(Ot) ~ V. Evidently at n M = 0. Because of the arbitrariness of the point t E X \ M this implies the openness of the set X \ M. Hence the set M is closed. The lemma is proved. • 8. Zorn's lemma and well ordered sets A relation -< between elements of a set M is called a partial order or partial ordering on the set M, if for some couples x, y E M we say that x precedes y (or y succeeds x) with respect to -< and we write x -< y (or y »- x), moreover: if x -< y and y -< z, then necessarily x -< Z; x -< y and y -< x if and only if x = y. In this case the set M is called partially ordered. We say that a set N ~ M is linearly ordered by the relation -< if for every two its elements x, y: either x -< y, or y -< x. An element x of a partially ordered set M is called maximal (resp., minimal) if there is no element of the set M, which succeeds (resp., which precedes) x and is not equal to x. A subset N of a partially ordered set M is called bounded if there exists an element of the set M which succeeds all elements of the set N. The following assertion is one of fundamental facts of mathematics, see [MEl or [En, §I.4]. Lemma 8.1 (Zorn). If every linearly ordered subset of a partially ordered set M is bounded then the set M contains a maximal element. • The relation of the inclusion ~ is an example of a partial order on the set of subsets. In the connection with this partial order a linearly ordered subset N ~ M is called maximal (linearly ordered subset of the set M) if it is not contained in another linearly ordered subset (different from N) of the set M. Lemma 8.2. Let N be a linearly ordered subset of a partially ordered set M. Then there exists a maximal linearly ordered subset Nl of set M containing N: N ~ Nl ~ M.
CHAPTER 1
34
Proof. Consider the set X of all linearly ordered subsets of the set M containing N and the partial order ~ on X. Every linearly ordered subset a of the set X is bounded (recall that we consider the partial order ~: the element Ua of the set X succeeds all elements of the set a). We obtain what was required from Zorn's lemma. A linearly ordered set M is called well ordered if every nonempty subset N of it possesses a minimal element (with respect to the order) minN E N. Example 8.1. The segment [0,1] of the real line with its natural order is not well ordered: its subset (0,1] does not possess a minimal element (0 tJ. (0,1]). Theorem 8.1 (Zermelo). Every set may be made well ordered. • Proof. Let X be an arbitrary set. Consider a family r of all couples (M,,), where M is a subset of X and, is a well ordering on M. Let us partially order the set r: we put (Ml"d -< (M2' '2) if Ml ~ M2; the order is a restriction to Ml of the order ,2 and Xl -< X2 with respect to the order ,2 for every Xl E Ml and X2 E M2 \ M l · The set r is nonempty: take as M an arbitrary one point subset of the set X; we need not define the order, in this case. By Lemma 8.2 there exists a maximal linearly ordered subset ~ of the set r. Let Xl = U{M: (M,,) E ~ for some ,}. Introduce the following linear order on the set Xl: if Xl, X2 E Xl then Xl, X2 E M for some (M,,) E ~, and we put Xl -< X2 if this is true with respect to the order,. As orders, for (M,,) E ~ are compatible our definition of the relation Xl -< X2 does not depend on the choice of (M,,) E ~. Show that the set Xl is well ordered by this relation. Let a set N ~ Xl be nonempty. Take an arbitrary Xl E N. Fix an arbitrary set M :7 Xl being the first member in a couple (M,,) E ~. The set Nl = N n M is nonempty. Since the set M is well ordered by " there exists Xo = min N l . Then Xo -< X for every X E Nl by virtue of the inclusion Nl ~ M and Xo -< X for every X E N \ M by virtue of our definition of the partial order on ~. Show that Xl = X. If this is not true then the set X \ Xl is nonempty. Take arbitrary Xl E X \ Xl' Make the set M = Xl U {xd well ordered (by an order ,) by putting X -< Xl for every X E Xl' On the one hand (M,,) E ~. On the other hand M g Xl' This contradiction gives what was required. The theorem is proved. • A one to one mapping
,1
,1
Topological and metric spaces.
35
case two isomorphic partially ordered sets may be related by several isomorphisms (for instance, for the real line every mapping ip.x(x) = AX with ). > 0 is an isomorphism). For well ordered sets this is not so. Assertion 8.1. Let ipl,ip2 : (M1,')'d ---t (M2 ,')'2) be two isomorphisms of well ordered sets. Then ipl == ip2· Proof. Let M = {x: x E M 1 , ipl(X) = ip2(X)}. The set Mis nonempty: it contains min M. Our aim is to prove that M = M 1 . If this is not so then the set N = Ml \ M is nonempty. Since the mappings ipl and ip2 are isomorphisms, we have ipl(minN) = ip2(minN). Then minN E M, which contradicts the definition of N. The assertion is proved. • Equivalence classes of well ordered sets with respect to the relation introduced of isomorphicity are called ordinal number.. Evidently the cardinalities of all specimens of every such class are equal. This gives the possibility of speaking about finite ordinal numbers. Here the situation is simpler than in the general case: if the cardinalities of two finite well ordered sets coincide then they are isomorphic. The reader may prove it as a simple exercise. Infinite ordinal numbers are called transfinite. Here the situation is more complicated. Example 8.2. The set of positive integers N and the set N U {oo} with their natural orders are well ordered and countable, but they are not isomorphic. We make ordinal numbers well ordered when we put a -< (3 if for some Ml E a and M2 E (3 and for some x E M2 there exists an isomorphism ip : Ml ---t {t: t E M 2 , t -< x} (evidently an analogous isomorphism ip and x exist for every specimens Ml E a and M2 E (3). Show that this order is linear, i.e., every two ordinal numbers a and (3 are comparable. Fix Ml E a and M2 E (3 and denote by P the set of all of elements s E Ml satisfying the condition:
(8.1) there exists an element t E M2 such that the well ordered sets Nl = {x x E M 1 , X -< s} and N2 = {y: Y E M 2 , Y -< t} are isomorphic. Assertion 8.1 implies the existence of the unique isomorphism of the set P onto an initial part Q of the set N 2 . Let ip be the corresponding isomorphism. If one of the sets Ml \ P or M2 \ Q is empty then we have what was required. In the opposite case we extend the isomorphism ip to the set PU{min(Ml \P)} by putting ip(min(Ml \P)) = min(M2 \Q). This means that min(Ml \ P) E P. This is impossible. We have what was required. Evidently because of this observation ordinal numbers become well ordered.
36
CHAPTER 1
Below we will use remarks in this section only in cases of an extreme necessity. These results do not give explicit descriptions of the orders mentioned. By Theorem 8.1 the segment X = [0,1] may be well ordered. The reader can try to describe such an order in an explicit form. But what follows from the possibility of making the set X well ordered? The set X is uncountable. Therefore if M is a countable well ordered set then for some t E X the set M isomorphic to the set {y: y E X, Y -< t}. This means that for some t E X the set n of all at most countable ordinal numbers is isomorphic to the set {y: y E X, Y -< t}. The ordinal number corresponding to the set of positive integers N is denoted by Wo. The ordinal number corresponding to the set n is denoted by WI. Evidently the set n is uncountable; see an analogous situation in Example 8.l. If a is an arbitrary element of a well ordered set M then the element a + 1 = min{ t: t >- a, t i= a} of the set M succeeds directly a. In this case the element a of the set M precedes directly the element a + 1 in the meaning that there is no elements of the set M lying between a and a + l. An element of a well ordered set is called isolated if a directly preceding element exists. The opposite situation is possible too. We mean for some element {J of the set M there is no element a, which directly precedes the element {J, i.e., for every a -< {J, a i= {J, there exists an element I i= a, {J of the set M lying between a and {J: a -< I -< {J. In this case the ordinal number (transfinite) {J is called limit. Remark 8.1. If a transfinite {J < WI is limit then it is the limit of a monotonically increasing (with respect to the order) sequence {JI -< {J2 -< {J3 -< ... : The set B = {a: a -< {J, a i= {J} is countable and we can enumerate its elements: B = {ai: i = 1,2, ... } (we forget temporarily the order -<). The set B = {al} U {ai: i = 2,3, ... , al, ... , ai-l -< ad is infinite. The order -< coincides on B with the order due to the numeration. After an obvious renumbering of its elements (with preservation of the order, see §1) we obtain the needed sequence {JI -< {J2 -< {J3 -< ....
CHAPTER 2
SOME PROPERTIES OF TOPOLOGICAL, METRIC AND EUCLIDEAN SPACES
In this chapter we continue our acquaintance with topological and geometric notions which will be used in the later account. 1. Tychonoff topology of products
Let {Xc>: a E A} be a family of topological spaces. In §1.1 we have defined a product of sets. Our nearest aim here is the introduction of a topology on the set X = IT{Xc>: a E A}. Notice first that we can do it in various ways. We will define a so called Tychonoff topology of the product, which corresponds best to our aims. Let Ao be an arbitrary finite subset of the index set A. For every a E Ao fix an open subset Uc> of the space Xc> and put
oc> =
{Uc> X c>
if a E A o, if a E A \ Ao.
The set IT {Oc>: a E A} is called a Tychonoff neighborhood (of each of its points). If Mc> and M~ are subsets of the set Xc> for every a E A then
Therefore the intersection of every pair of Tychonoff neighborhoods is a Tychonoff neighborhood also. The set X is also a Tychonoff neighborhood. So the set of all Tychonoff neighborhoods satisfies conditions A and B of §1.2. Hence it is a base of an uniquely defined topology on the set X. The topology is called the Tychonofftopology of the product IT{Xc>: a E A}. It is easier to understand geometric particularities of a topological space when we have a simply described base of the space or bases at its points. In this connection let us notice that if for every a E A a base f3c> of the space Xc> (respectively, a base f3c> at a point Xc> E Xc» is fixed, then the family of all Tychonoff neighborhoods satisfying (with the above notation)
38
CHAPTER 2
the condition Uo. E f3o. is a base of the space X (respectively, a base at the point x = {xo.: a E A} EX). For every a E A let Yo. be a (nonempty) subspace the space Xo.. It is natural to consider the product Y = IT {Yo.: a E A} as a subset of the product X. So we have two topologies on the set Y. The first is the topology 71 which is induced on Y by the topology of the space X. The second is the Tychonoff topology 72 of the product IT {Yo.: a E A}. The formula (1.1) and the definition of the induced topology imply that if 0 is a Tychonoff neighborhood of the product X then the set 0 n Y is a Tychonoff neighborhood of the product Y. Every Tychonoff neighborhood in Y may be represented in the form of such an intersection. So the same base generates on Y both the topology 71 and the topology 72. Hence these topologies coincide. Let B be a subset of the index set A. The product X B = IT {Xo. : a E B} is called subproduct of the product IT {X a: a E A}. If the set B consists of one element a usually we speak not about a subproduct but about a factor Xo.. We may regard the subproduct X B as 'embedded' in the product. To define an 'embedding' for every a E A \ B we fix a point x~ E Xo.. For a point y = {Yo.: a E B} E X B we now put i(y) = {xo.: a E A}, where Xo. = Yo. for a E Band Xo. = x~ for a E A \ B. Our definition of an 'embedding' is not unique. It depends on the choice of the points x~ for a E A \ B. The mapping PB : X ----+ X B , PB({X",: a E A}) = {x",: a E B}, is called a projection. The projection is defined uniquely. For the projection of the product X = IT {X",: a E A} on the factor X{3 the preimage of an open subset U of the factor X{3 has the form of a Tychonoff neighborhood IT {O",: a E A}, where 0{3 = U and 00. = Xo. for a E A \ {f3}. Therefore Theorem 1.7.5b implies
Lemma 1.1. A projection of a product onto a factor is continuous . • Lemma 1.2. LetX = IT{X{3: f3 E B} be a product of topological spaces X{3, f3 E B, {xo. = {xo.{3: f3 E B}: a E A} be a generalized sequence of points of the space X. Let for every f3 E B the generalized sequence {xo.{3 : a E A} con verge with respect to a directed set
Some properties of topological, metric and Euclidean spaces.
39
mapping f is continuous if and only if for every a E A the 'coordinate' mapping fa = Paf is continuous. Proof. Necessity follows from Lemma 1.1 and from the fact that the composition of continuous mappings is continuous too. Sufficiency follows from Lemma 1.2. The theorem is proved. • When we compare Lemma 1.1 and Theorem 1.1, we obtain Corollary. A projection of a product of topological spaces onto every is subproduct is continuous. • It is natural to describe a mapping into a product by the definition of coordinate mappings (we do so, for instance, when we introduce a vector function with values in the three-dimensional space by writing separate formulae for the dependence on the argument of X-, Y- and z-coordinates). The inverse passage is possible too: Assume that for a E A a mapping fa : X --t Y a is fixed. The mapping f : X --t IT {Ya: a E A}, f(x) = {fa (X): a E A} is called a product (or a diagonal product) of the family {fa: a E A}. Theorem 1.1 implies: a diagonal product of continuous mappings is also a continuous mapping. Now let factors in the product be metric spaces. Does this cause any particularity in the questions under consideration? In particular, whether Tychonoff topology of the product is generated by a metric related in a natural way with metrics on factors? We will give an affirmative answer to the last question under some additional assumptions. Let diam M denote the diameter sup{p( s, t): s, t E M} of a subset M of a metric space (X, p): diam M = sup{p(s, t): s, t EM}. Consider the following situation:
(1.2) (Xb Pk), k
= 1,2, ... , is a sequence of metric spaces and diamXk
--t
O.
Under these assumptions define the distance d on the set X = IT {X k : k = 1,2, ... }. For x = {Xk: k = 1,2, ... } and Y = {Yk: k = 1,2, ... } put d(x, y) = SUP{Pk(Xk, yd: k = 1,2, ... }. Our assumptions imply that only a finite number of elements of the sequence {PdXk' yd: k = 1,2, ... } may be greater than 1. Therefore values of the function d are finite. The fulfilment of conditions 1 and 2 of the definition of a metric in §1.4 is obvious. Check the fulfilment of the triangle inequality. Let in addition Z = {Zk: k = 1,2, ... }. For every k = 1,2, ... Pk(Xk, Zk) ~ Pk(Xk, Yk)
+ pdYk, Zk)
~ d(x, y)
+ d(y, z).
Since the last expression does not depend on k, d(x, z)
= SUP{Pk(Xk, zd: k = 1,2, ... } ~ d(x, y) + d(y, z).
Which gives us what was required.
40
CHAPTER 2
Theorem 1.2. Under the condition (1.2) the Tychonoff topology of the product X = IT {X k: k = 1, 2, ... } is generated by the above metric d. Proof. I. Let a set U ~ X be open in the Tychonoff topology and x = {Xk: k = 1,2, ... } E U. By the definition of the Tychonoff topology there exists a number i = 1,2, ... and for every k = 1, ... , i there exists a neighborhood Vk of the point Xk in the space X k such that
I1{Vk : k=l, ... ,i}xI1{Xk : k=i+1,i+2, ... }~U. For every k = 1,2, ... , i fix a number [(k) > 0 such that OE(k)Xk ~ Vk. Put [ = min{E(l), ... , [(i)}. It remains to notice that
O"X ~ I1{O"Xk: k = 1, ... ,i} x I1{X k : k = i + 1,i + 2, ... } ~I1{Vk: k=l, ... ,i}xI1{Xk : k=i+1,i+2, ... }~U. In view of the arbitrariness of the point x E U the last fact means the openness of the set U with respect to the metric d. II. Take an arbitrary point x = {Xk: k = 1,2, ... } E X and an arbitrary number [ > O. Find a number i such that diam X k < [ for every k = i + 1, i + 2, .... For k = i + 1, i + 2, ... we have OcXk = X k. By virtue of our definition of the metric d on the product
O"X=I1{O"Xk: k=l, ... ,i}xI1{Xk : k=i+1,i+2, ... }, i.e., [-neighborhood of a point with respect to the metric d has the form of a Tychonoff neighborhood. So if a subset of the product is open with respect to the metric d, then it is open with respect to Tychonoff topology too. When we compare I and II we obtain that the Tychonoff topology on X coincides with the topology generated by the metric d. The theorem is proved. • In the next section we will show that the additional condition diam X k ---7 o for the existence of some metric on the product is inessential, but the above concrete description of the metric on the product will not appropriate. Every finite family of spaces may be extended to a countable family by taking as missing elements a one point space (a metric on it can be defined in a unique way). These additional elements change nothing in the consideration of the Tychonoff topology as well as in the introduction of the metric d. The condition diamX k ---7 0 is obviously fulfilled for the family completed in this way. Therefore the theorem proved may be automatically restated to a finite number of factors. In addition, we can propose to the reader as a simple exercise to revise the given reasoning, to reject conditions
Some properties of topological, metric and Euclidean spaces.
41
and remarks related with the infiniteness of the number of factors, and to give (an easier) direct proof of the existence of a metric on the product of a finite family of metric spaces. Example 1.1. The metric on the space JRn in Example 1.4.5 coincides with the metric introduced on this set as on the product of lines according to Theorem 1.2. Example 1.2. Let continuous real functions f and 9 be defined on a topological space X. By Theorem 1.1 the mapping
: X --+ JR x JR, (x) = (f(x),g(x)), is continuous. Since the composition of continuous mappings is continuous, the remarks of Examples 1.7.2 and 1.1 imply the continuity of the function h( x) = f (x) + g( x). Example 1.3. The remarks of Example 1.7.4 and Theorem 1.7.7 imply the continuity of the function cp : JR --+ JR, cp( t) = e. Since
the remarks of the previous example imply the continuity of the mapping 'lj; : JR x JR --+ JR, 'lj;(x, y) = xy. The reasoning of the previous example now allows us to prove that under the same assumptions the function hI (x) = f(x)g(x) is continuous. Thus if a function can be constructed with help of the operations of addition and multiplication of continuous functions, then it is continuous. This remark and Lemma 1.1 imply the continuity of every mapping \]I : JRm --+ JRn with polynomials as coordinate functions (we consider the spaces with the metric of Example 1.4.5). Theorem 1.3. Let the hypotheses of Theorem 1.2 hold. Let the spaces Xb k = 1,2, ... , be compact. Then the product X is compact. Proof. Since by the previous theorem the space X is metric, to prove its compactness it is sufficient to check the fulfilment of condition (1.6.3). Let I be an arbitrary sequence of elements of the space X. We will choose sequentially subsequences of the sequence I' First we put 10 = I' Let the sequence Ik-I be constructed. Our next step is the construction of a sequence Ik. We do it as follows. Consider the sequence a of k-th coordinates of elements of Ik-I' Since the space X k is compact the sequence a contains a subsequence al converging to a point Yk E X k • We take as Ik the subsequence of the sequence Ik-I consisting of elements of Ik-I' the k-th coordinates of which belong to the sequence al' Now construct a subsequence 1* = {Xi; : j = 1,2, ... } of the sequence I' As Xit we take an arbitrary element of the sequence 11' Let elements Xi t , . . . ,Xi;_t be chosen. As Xi; we take an arbitrary element of the sequence Ij with the index > i j - I . Elements of 1* with indices > k belong to the sequence Ik' Therefore the k-th coordinates of the elements of 1* constitute
CHAPTER 2
42
a sequence converging to the point Yk. By Lemma 1.2 the sequence --y* converges to the point Y = {Yk: k = 1,2, ... }. The theorem is proved. • Theorem 1.4. Let X be a compactum and Y be an arbitrary topological space. Then the mapping F : Y -+ X x Y, F (y) = X x {y}, is upper semicontinuous. Proof. The proof use Theorem 1. 7.5. Let H be an arbitrary closed subset of the product X x Y and Y E Y\F- 1 (H). The last fact means that the intersection (X x {y} ) n H is empty: (X x {y }) n H = 0. Therefore for every x E X we can fix a Tychonoff neighborhood Ox x Vx of the point (x, y) in the product X x Y which does not meet the set H. The cover {Ox: x E X} of the compactum X contains a finite subcover {OX1,"" Oxd. Put Oy = n {Vx i l . . . , VXk } . Every point x E X belongs (at least) to one of the sets OXi' Therefore {x} x Oy <;;; OXi X VXi <;;; (X X Y) \ H. In view of the arbitrariness of the point x E X we have:
(X x Oy) n H = 0,
Oy n F- 1(H) =
0,
y ~ [F-l(H)].
Since the point y E Y \ F- 1 (H) was taken arbitrarily, the last fact means the closedness of the set F- 1 (H). The theorem is proved. • Corollary. The product of two compacta is a compactum. To obtain the corollary it is sufficient to remember Theorem 1.7.8. •
2. Continuous bijections. Homeomorphisms. Metrizability A continuous bijection is called an homeomorphism if its inverse mapping is continuous too. Let f : X -+ Y be an homeomorphism of a topological space X onto a topological space Y. We have one to one correspondence between elements of these two sets. Under these assumptions a set M <;;; X is open in the space X if (by virtue of the continuity of the mapping f) and only if (by virtue of the continuity of the mapping f-l) the set f(M) is open in the space Y (see Theorem 1.7.5). In other words, our one to one correspondence is extended to topologies of these spaces (with preservation of all relations of inclusion and belonging). This means that every fact which is stated in these terms and which is true for one of these homeomorphic spaces, is true for the other too. So the existence of an homeomorphism means a topological equivalence, an identity of topological properties of the spaces. For instance, if one of these spaces is Hausdorff, then the second one is Hausdorff too. If one of these spaces is compact, then the second one is compact too, etc .. We say that a space X homeomorphic to a space Y if there exists an homeomorphism of the space X onto the space Y. An identity mapping is an homeomorphism. Thus every space is homeomorphic to itself. Since a mapping that is inverse of an homeomorphism is an homeomorphism also,
Some properties of topological, metric and Euclidean spaces.
43
the relation of homeomorphism is symmetric: if a space X is homeomorphic to a space Y then the space Y is homeomorphic to the space X. Therefore usually we say more briefly that the spaces X and Yare homeomorphic. The composition of two homeomorphisms is an homeomorphism. Therefore (in view of the previous remark) if two spaces are homeomorphic to a third then they are homeomorphic to each other. A mapping f : X ---7 Y is called an embedding if when we consider it as a mapping of the space X onto the space f(X) it become an homeomorphism. This word (in inverted commas) was used in the previous section.
Example 2.1. Let real numbers a and b be different. Remarks at the end of the previous section imply the continuity of the function f(t) = (1 - t)a + tb and of its inverse function g(x) = (x - a)j(b - a). The segment [0, 1] is mapped by the function f onto the segment I with endpoints a and b (i.e., I = [a, b] if a < b, and I = [b, a] if a > b). The point goes into the point a and the point 1 goes into the point b. The mapping of the segment [0, 1] and its inverse mapping are the restrictions of the continuous mappings f and g. Thus they are continuous. They are homeomorphisms. We have proved: every (nondegenerate) segment of the real line is homeomorphic to the segment [0,1]. Corollary: Any two segments of the real line are homeomorphic.
°
Example 2.2. Every pair of intervals of the real line are homeomorphic. The proof for intervals with finite endpoints is analogous to the reasoning of the previous example. The function tan is an homeomorphism of the interval ( - ~, ~) onto the line ~ (the inverse function arctan is continuous too). The function f (t) = a + e t realizes an homeomorphism of the line onto the interval (a, 00). The function g(t) = a-e t realizes an homeomorphism ofthe line onto the interval (-00, a) (the inverse functions h±(t) = In(±(x - a)) are continuous too). Example 2.3. A segment is not homeomorphic to an interval: A segment is compact and an interval is not compact. Example 2.4. Affine transformations are continuous (see the end of the previous section). So affinely equivalent subspaces of the plane are homeomorphic. Thus any two triangles are homeomorphic, and the boundaries of any two triangles are homeomorphic. Example 2.5. Not every continuous bijection is an homeomorphism. Let (X, T) be an arbitrary topological space. Apart from the topology T let us consider the discrete topology Td on the set X. The identity mapping of the space (X, Td) onto the space (X, T) is continuous. This follows immediately from the definitions. It is an homeomorphism (i.e., its inverse mapping is continuous too) if and only if the topology T is discrete. Since not every topology is discrete, we obtain the required example.
44
CHAPTER 2
Theorem 2.1. A continuous bijection of a compact space onto a Hausdorff space is an homeomorphism. Proof. The proof reduces to referring to Corollary of Theorem 1. 7.S .• Often the presence of additional structures simplifies the investigation of topological spaces. We may see this in the previous account where we studied both a metric and a topology. Notice that in corresponding results as a rule we made conclusions about topological properties of spaces under consideration, i.e., the metric was not mentioned in the description of these properties. In such situations properties of any particular metric were not important and only the existence of a metric generating the topology in question played any role. A topological space (X, r) is called metrizable if there exists a metric on the set X generating the topology r. If there exists a one to one correspondence f between a set X and a metric space (Y, p), then we can define a metric d on the set X by putting d(s, t) = p(f(s), f(t)) for s, t EX. If in addition f is an homeomorphism of a topological space (X, r) onto the metric space (Y, p) then the metric d generates the topology r. Therefore a topological space which is homeomorphic to a metric space is metrizable. The inverse fact is obvious: a metrizable space is homeomorphic to a metric space. We speak here not about the existence of some metric on the given set. Such a metric always exists (see Examples 1.4.3 and 1.4.9, where a metric generating the discrete topology is pointed out). In Example 2.5 we have shown even that every space is the image of a discrete space under a continuous bijection. We discuss here the question of when, for a given topology, we can find a metric generating just this topology. So the discrete topology is metrizable (see Examples 1.4.3 and 1.4.9), although the initial definition of the discrete topology is not related to any metric. Example 2.6. The space in Example 1.6.3 is metrizable if and only if it is (at most) countable. In a metric space every point has an (at most) countable base, for instance the base consisting of its i-neighborhoods, k = 1,2, .... Thus every point may be represented as the intersection of an (at most) countable family of open sets. In our case the complement of every neighborhood of the point a (the notation of Example 1.6.3) is finite. Therefore the point a may be represented as an intersection of a countable number of open sets only in the case when the space X is countable. Consider the case of a countable space X. Enumerate elements of the set A : A = {ak: k = 1,2, ... }. Define the mapping f of the space X into the real line by putting f(a) = 0 and f(ad = for k = 1,2, .... The continuity of the mapping f follows easily from the definition of the topology of the space X. Evidently the mapping is injective. By Theorem 2.1 the space is homeomorphic to the subspace f(X) of the real line. The case of a finite space X is trivial. Such a space is discrete. Remark 2.1. Let for metrics PI and P2 on a set X there exist positive
i
Some properties of topological, metric and Euclidean spaces. numbers a and
!3
45
such that for every s, t E X
Then the topologies generated by the metrics PI and P2 coincide. In fact, the condition mentioned may be rewritten in the form of the system P2 ( s, t) ~ ~PI (s, t) { PI (s, t) ~ - P2 (s, t).
a
The first inequality of the system means the fulfilment of the Lipschitz condition for the identity mapping (X, PI) - t (X, P2). The second one means the fulfilment of the Lipschitz condition for the inverse mapping. Thus the identity mapping is an homeomorphism, as was required. In Example 1.4.2 we have introduced the Euclidean metric P in the space ]Rn. In Example 1.4.5 we have introduced on the same set the metric d. By virtue of Theorem 1.2 the metric d generates the Tychonoff topology in the space ]Rn when we consider it as the product of lines. These metrics are related by the inequality (2.1)
d(s, t) ~ p(s, t) ~ vnd(s, t).
Remark 2.1 implies the coincidence of the generated topologies, i.e.: Remark 2.2. The Euclidean topology on ]Rn coincides with the topology of the Tychonoff product. Remark 2.3. Let (X, PI) be a metric space, a > o. For s, t E X put P2(S,t) = min{PI(s,t),a}. Evidently the function P2 is a metric. The metrics PI and P2 generate the same topology on the set X: the family {{s: SEX, PI(S,t) < c}: 0 < c < a} is a base at the point t E X both for the topologies of the space (X, PI) and of the space (X, P2). By this remark the condition diamX k - t 0 in Theorems 1.2 and 1.3 is not very essential, i.e., we can change the metrics on the factors in a manner that the condition turns out to be fulfilled and the topologies are not changed. Therefore Theorem 1.2 implies: Theorem 2.2. A product of an {at most} countable family of metric spaces is metrizable. • Theorem 1.3 implies Theorem 2.3. A product of an {at most} countable family of metric compacta is a metric compact. • Example 2.7 'Hilbert cube'. Hilbert cube is the countable power of the segment. By Theorem 2.3 it is a metric compactum. Example 2.8. The space D consisting of the two points 0 and 1 (with the discrete topology) is called a two-point space. By Theorem 2.3
CHAPTER 2
46
its countable power DN is compact. Show that the space DN is homeomorphic to the Cantor perfect set. Use the description of the Cantor perfect set in Example 1.6.4. For k = 1,2, ... define the function ik in the following way. Let Sk-l([O,l]) be the union of disjoint segments [ai, bi ], i = 1, ... ,2 k- 1 (SO([O,l]) = [0,1]). Put the function fk equal to 0 on every set K n [ai, ai + t(b i - ai)] and equal to 1 on every set K n [ai + }(bi - a;), bd, i = 1, ... ,2 k - l . Each of these sets is closed in K. The function fk is constant on it. Hence the function is continuous there. Theorem 1.7.6 now implies the continuity of the function ik. Theorem 1.1 implies the continuity of the diagonal product f of mappings fk' k = 1,2, .... We propose that the reader, as a simple exercise, shows that the pre image of every point of the space K under the mapping f consists just of one point. Referring to Theorem 2.1 completes the reasoning. 3. Some properties of a metric Lemma 3.1. Let A be a nonempty subset of a metric space (X,p). Then Ip(x, A) - p(y, A)I ::::; p(x, y) for every points x, y of the space X Proof. I. Prove first that
p(p, A) ::::; p(p, q)
+ p(q, A)
for every points p, q EX. For tEA we have p(p, A) ::::; p(p, t) ::::; p(p, q) + p(q, t). When we pass to the lower bound with respect to tEA in the right hand side of this inequality we obtain the estimate needed. II. If we substitute in the inequality of! first p = x, q = y and next p = y, q = x, we obtain p(x, A) - p(y, A) ::::; p(x, y), p(y, A) - p(x, A) ::::; p(x, y). This gives what was required. • Corollary 1. Let (X, p) be a metric space. Let Sl, 82, t l , t2 EX. Then
because
+ (p(s2,td - P(S2,t2))1 P(S2' tdl + Ip(S2' td - P(S2' t 2)1
Ip(Sl,td - P(S2,t 2)1 = I(P(Sl,td - P(S2,t l )) ::::; Ip(Sl' t l ) ::::; P(Sl' S2) + P(tl' t2)'
Corollary 2. Let A be a nonempty subset of a metric space (X, p). Then the function cp(t) = p(t, A) is continuous on X. Because by Lemma 3.1 the function cp satisfies the Lipschitz condition with the constant 1. •
Some properties of topological, metric and Euclidean spaces.
47
Lemma 3.2. With the notation of Corollary 2 of Lemma 3.1 the set A is closed in X if and only if A = '11-1(0). Proof. The proof is obvious (see Remark 1.4.1). • By analogy with c-neighborhoods of points: for c > 0 and a nonempty subset M of a metric space (X, p) the set OoM = O(M,c) = {t: t E X, p(t,M) < c} is called c-neighborhood of the set M. Let also O,(M,c) = {t: t E X, p(t,M) ~ c} for c ~ O. Corollary 2 of Lemma 3.1 implies immediately: Lemma 3.3. With the above notation the set O(M, c) is open and the set O,(M, c) is closed in the space X. • The definition implies: Lemma 3.4. A point x of a metric space (X, p) belongs to the c-neighborhood of a nonempty subset A of the space X if and only if p(x,t) < c for a point tEA. In other words Oo(A) = U{Oot: tEA}. • Lemma 3.4 and the triangle inequality imply: Lemma 3.5. Let A and B be nonempty subsets of a metric space X. Let a and f3 be positive numbers and B ~ Oo:A. Then O{3B ~ Oo:+!3A. • Lemma 3.6. Let A and B be nonempty subsets of a metric space X and the set B be compact. Then there exists a positive number a such that B ~ Oo:A. Proof. The family, = {OoA: c > O} is an open cover of the space X. By virtue of the compactness of the set B the family , contains a finite subfamily {OOI A, ... , OOk A}, which cover B too. Evidently a = max {Cl, ... , Ck} satisfies the above conditions. The lemma is proved .• A subset M of a metric space X is called bounded if there are a point x E X and a number c > 0 such that M ~ Oox. Remark 3.1. The diameter of every bounded set is finite. Remark 3.2. If M is a subset of finite diameter of a metric space X and x is an arbitrary point of the space X, then M ~ Oox for some c > o. Proof of these assertions reduces to a simple verification with the help of the triangle inequality. They imply, in particular, that the conditions of the boundedness of the set and of the finiteness of its diameter are equipotent. On the other hand if we take a one point set as A in Lemma 3.6 we obtain: Lemma 3.7. Every compact subset of a metric space is bounded. • Lemma 3.5 implies: Remark 3.3. c-neighborhood (c > 0) of a bounded set is a bounded set too. Assertions of §1.6 and Lemma 3.7 imply: Remark 3.4. Let a real function f be defined and be continuous on a compact space X. Then -00 < inff(X) ~ supf(X) < 00 and {inff(X), sup f(X)} ~ f(X).
48
CHAPTER 2
As a particular case of Remark 3.4 we obtain: Remark 3.5. Let a positive function f be defined and be continuous on a compact space X. Then < inf f(X) ~ supf(X) < 00 and {inf f(X), sup f(X)} ~ f(X). A distance between nonclosed disjoint (nonempty) sets may be equal to zero. This is sufficiently obvious. We have already noticed that if a point x belongs to the closure of a set M, then p(x, M) = 0. The distance between two closed disjoint sets may be zero too, see below Example 4.2. However: Lemma 3.8. Let A and B be nonempty closed disjoint subsets of a metric space (X, p). Let the set B be compact. Then p(A, B) > and there exists a point b E B such that p(A, b) = p(A, B). Proof consists in referring to Corollary of Lemma 3.1 (with the notation of Lemma 3.1) and to Remark 3.5 (with f = 0. In other words, [-neighborhoods of a compact subset of a metric space constitute a base of its neighborhoods. • A topological space is called normal if it is Hausdorff and every pair of disjoint closed subsets of it possess disjoint neighborhoods. Evidently every normal space is regular. Lemma 3.9. Every metric space is normal. Proof. Let F1 and F2 be closed disjoint subsets of a metric space (X, p). The function g : X ~ ]R,
°
°
p(t, Fd - p(t, F 2 ) g () t = '-'------'----'--'------'p(t, Fd + p(t, F2)'
is continuous (see Corollary 2 of Lemma 3.1 and §1). The sets OF1 = g-1(( -00,0)) and OF2 = g-1((0, (0)) are disjoint neighborhoods of the sets F1 and F 2, respectively (see Theorem 1.7.5).
The lemma is proved.
•
4. Some properties of Euclidean, locally compact, and normed spaces
The metric of the Euclidean space ]Rn is defined in Example 1.4.2, see also Remark 2.2 and Example 1.1. Lemma 4.1. The set 0,(0, r), where r > 0, is compact. Proof. The inclusion 0,(0, r) ~ [-r, r]n (~ ]Rn) is obvious. The set [-r, r]n is compact by virtue of the compactness of the segment [-r, r] and Theorem 1.3. Lemma 3.3 and Theorem 1.6.2 imply what was required. The lemma is proved. •
Some properties of topological, metric and Euclidean spaces.
49
Theorem 4.1. A subset of an Euclidean space is compact if and only if it is closed and bounded at once. Proof. Sufficiency follows from remarks of the previous section, Lemma 4.1 and Theorem 1.6.2. Necessity follows from Theorem 1.6.1 and Lemma 3.7. The theorem is proved. • This theorem immediately implies that the closure of every bounded subset of an Euclidean space is compact, because this closure is a bounded set too. A topological space is called locally compact if every point possesses a neighborhood the closure of which (in the space under consideration) is compact. Evidently every compact space is locally compact. Every open subspace of a compactum is locally compact. The remark following Theorem 4.1 implies the local compactness of open subsets of the Euclidean space. So the following assertion is applicable directly to an arbitrary open subset U of the Euclidean space. Lemma 4.2. Let U be an arbitrary neighborhood of a nonempty compact subset B of a locally compact metric space X. Then there exists a number b > 0 such that the set 0f(B, b) (~O(B, b)) is compact and lies in U. Proof. Show first that there exists a neighborhood V of the set B such that the set [V] is compact and lies in U. According to the local compactness of the space X associate to every point b of the set B its neighborhood Wb such that the set [Wb ] is compact. According to the normalness of the space X (see Lemma 3.9) associate to each point b E X of the set Bits neighborhood Vb and a neighborhood Nb of the set Fb = X \ (Wb n U) such that Vb n Nb = 0. Then [Vb] n Fb = 0, [Vb] S;;; un W b. By virtue of the inclusion [Vb] S;;; Wb the set Vb is compact. The cover {Vb: b E B} of the compactum B has a finite subcover {Vb" ... , Vb.}. The set V = Vb, U·· ·UVb• lies in the compactum
that gives what was required. Our assertion now follows from Corollary of Lemma 3.8 (with V as the set U of the corollary) and from Lemma 3.3 .• Later we will use: Assertion 4.1. Let X be a locally compact metric space and M be its (at most) countable nonempty closed subset. Then set M has an isolated (in M) point. Proof. Take an arbitrary point p E M. By virtue of the local compactness of the space X for some c > 0 the set 0f(P,c) is compact. The set G = O(p, c) n M is open in the compactum K = O,(p, c) n M. It is (at most) countable. Enumerate its elements: G = {Xk: k = 1,2, ... }. Apply Baire's theorem 1.6.7 to the open subset G of the compact urn K. At least
CHAPTER 2
50
one of the sets {xd, k = 1,2, ... , has a nonempty interior. Let {xd be such a set. Since the set {xd consists of one point, the nonemptiness of its interior means that ({ xd) = {xd. Hence the set {xd is open (in K). Thus the point Xi is isolated (in G ~ K and hence it is isolated in M). The assertion is proved. • Corollary. Let U be an open subset of the space IRn. Let M be an (at most) countable nonempty closed (in U) subset of the set U. Then the set M has an isolated point (in M). • Let X and y be two arbitrary points of a vector space L. The set {tx + (1 - t)y: 0 ~ t ~ I} is called a segment (connecting the points x and y. The points x and yare called endpoints of the segment). A subset M of the space L is called convex, if for every two its points the set M contains the segment connecting the points. In order to investigate and to use properties of convex sets we will need the following notion. A function f : L ----t IR is called a linear functional, if f(u + v) = f(u) + f(v) and f(au) = af(u) for every elements u, v of the space L and for every number a. A very simple example of a linear functional on the space IR n is the function ai, i = 1, ... ,n, which associates to an element x = (Xl,' .. , x n ) of the space IRn its i-th coordinate Xi' This example may be generalized in the following way. Fix an arbitrary vector e = (al,"" an) E IRn. We define the linear functional
alxl
+ ... + anx n·
In fact, every linear functional f : IRn ----t IR may be represented in such a form. For i = 1, ... ,n let aj = f(O, . .. ,0,1,0, ... ,0), where the unity is at the i-th position. We obtain
This implies, in particular, the continuity of the functional f, i.e., every linear functional on the space IRn is continuous. When we introduce the notion of a metric space we take as a pattern subsets of the line, of the plane and of three-dimensional space. However not all properties of such sets have analogs in an arbitrary metric space. We are closer to the usual geometry when instead of a metric we consider a norm. A real function which is defined on a vector space L and which associates to a vector u E L a number denoted by lIull, is called a norm, if: 1) Ilull > 0 for every nonzero vector u E L; 2) Ilaull = lal . Ilull for every vector u ELand every number a; 3) Ilu + vii ~ lIull + Ilvll for every two vectors u, vEL. The number lIull is called the norm of the vector u. A pair consisting of a vector space and a norm on the space is called a normed space. Length of vectors in Euclidean space is a norm, i.e., it satisfies the listed conditions.
Some properties of topological, metric and Euclidean spaces.
51
In a normed space we introduce a metric as follows. For u, vEL we put p(u, v) = Ilu - vii. The verification that the function p is a metric is easy. Thus every normed space may be considered as a metric space. We can apply to normed spaces the terminology and theory elaborated for metric spaces. Lemma 4.3. Let M be a (nonempty) convex subset of a normed space. Then for every
f
>
°
the set OEM is convex.
Proof. Let p and q be arbitrary points of the set OeM. By Lemma 3.4 there are points x and y of the set M such that Ilx - pil < f and Ily - qll < f. Take now an arbitrary point a of the segment connecting the points p and q. The point a may be represented in the form tp + (1 - t)q, t E [0,1]. By virtue of the convexity of the set M the point b = tx + (1 - t)y belongs to the set M. On the other hand,
Iia - bll
+ (1 - t)(q - y)11 xii + (1 - t)llq - yll
= Ilt(p - x) ~ tllp -
< tf + (1 =
t)f
f.
Hence a E OEM. The lemma is proved. • Lemma 4.4. The intersection of every family of convex sets is a convex set.
Proof. The proof is obvious: A segment connecting two arbitrary points of the intersection lies in every element of the family. Therefore it lies in the intersection. The lemma is proved. • Lemma 4.5. The closure of a convex subset of a normed space is a convex set.
Proof. Let M be a (nonempty) convex subset of the normed space under consideration. The representation [M] =
52
CHAPTER 2
points of the stated form is convex. This may be easily proved by a direct verification. Thus the convex hull of a set M coincides with the set of all points of the form alxl + ... akXk, where Xl, ... ,Xk E M, al,· .. ak are non-negative number and al + ... + ak = 1. (Such linear combinations are called convex combinations.) Evidently the convex hull of a convex set coincides with the set. By virtue of Lemma 4.5 and the equality cc(M) = [c(M)] its closed convex hull coincides with its closure. Let A and B be subsets of a metric space (X, p). A point a E A is said to be nearest to the set B, if p(a, B) = p(A, B). The point b of the set B in Lemma 3.8 is nearest to the set A. Notice that in the definition we do not require the uniqueness of a nearest point. Lemma 4.6. Let A and B be nonempty closed subsets of the space JRn. Let the set B be compact. Then the set A has a point that is nearest to the set B. Proof. Take an arbitrary number c > p(A, B). The distance from points of the set A \ OcB to the set B is not less than c. Since p(A, B) < c, p(A, B) = p(A n [OcB], B). The set An [OcB] is compact. By Lemma 3.8 there exists a point a E An [OcB] such that p(a, B)
= p(A n [OcB], B) = p(A, B).
The lemma is proved. • Lemma 4.7. Let a segment I connect points p and q in the space JRn. Let X E JRn. Then the function r.p: I ~ JR, r.p(t) = p(t,x), takes its maximal value M at p or at q (or both at p and at q) and values of the function r.p at other points of the segment I are strongly less than M. Proof. Let 0 :::;; s :::;; 1 and t = sp + (1 - s)q = q + s(p - q). Then r.p(t) =
lit - xii
= V(q - x
= V(t - x, t - x)
+ s(p -
= V(q - x, q - x)
q), q - x
+ 2s(q -
+ s(p -
q))
X,p - q)
+ S2(p -
q,p - q).
The last subradical expression as a function on s has the form of a square trinomial. The coefficient in S2 is positive. Therefore the maximum of this expression occurs at one of the endpoints (or at both endpoints) of segment [0,1], i.e., for s = 0 or 1. Respectively, the maximal value of the function r.p occurs at t = P or at q. The lemma is proved. • Corollary. Let A be a non empty closed convex subset of the space JRn and x E ~n. Then there exists a unique point of set A being nearest to the point x.
Some properties of topological, metric and Euclidean spaces.
53
The existence of such a point is proved in Lemma 4.6. Let us prove its uniqueness. Assume that the set A has two different points p and q nearest to the point x. By Lemma 4.7 the segment connecting the points p and q (it lies in the set A) has a point that is nearer to the point x than the points p and q. This is impossible by the definition of a nearest point. • Example 4.1. Consider the segments A = {O}x[O, 1] and B = {1} x [0, 1] of the plane 1R2 (Figure 2.1). Every point of the set A is nearest to the set B and every point of the set B is nearest to the set A.
B
A
A Figure 2.1
Figure 2.2
Example 4.2. Consider the x-axis A and the set B = {(x, y) : x > 0, y ~ ~} (Figure 2.2) of the plane 1R2. Both of them are closed and convex. The distance between the sets A and B is equal to zero. Since they are disjoint the set A has no nearest point to the set B and the set B has no nearest point to the set A. Theorem 4.2. Let A and B be nonempty closed convex disjoint subsets of the space IRn. Let the set B be compact. Then there exist a linear functional f : IR n --+ IR and real numbers p < q such that f(A) ~ (-oo,p] and f(B) ~ [q, (0). Proof. Let a point a E A be nearest to the set B. Let a point b E B nearest to a (the existence of such points is guaranteed by Lemma 4.6). For u E IR n put f(u) = (u, b - a). 1. Let x E A. The point a E A is nearest to the set B. The segment with the endpoints a and x lies in A. Thus for s E (0,1]
II(sx + (1 ((a (a - b, a -
bll ~ p(sx + (1 - s)a, B) ~ p(a, B) = Iia - bll, II(a - b) + s(x - a)11 ~ Iia - bll, b) + s(x - a), (a - b) + s(x - a)) ~ (a - b, a - b), b) + 2s(a - b, x - a) + S2(X - a, x - a) ~ (a - b, a - b), 2(a - b, x - a) + s(x - a, x - a) ~ O.
- s)a) -
As it is true for every s E (0,1], then (a - b, x - a) f(x) ~ f(a).
~
0, (x, b- a)
~
(a, b- a),
CHAPTER 2
54
II. Let y E B. The point b E B is nearest to the point a. The segment with the endpoints band y lies in B. As in I for s E (0,1]
II(b 2(y - b, b -
+ s(y - b) II = IIsy + (1 a) + s(y - b, y - b) ~ 0, (y a)
s)b - all ~
lib -
all,
b, b - a) ~ 0, f(y) ~ f(b).
III. Since (b - a, b - a) > 0, f(b) > f(a). Put p = f(a) and q = f(b). By I and II all required conditions hold. The theorem is proved. • The just proved Theorem 4.2 is also a consequence of the Hahn-Banach theorem, see below §5.2. 5. Remarks on multi-valued mappings
Lemma 5.1. Let
final to
°
space X into a metric space Y and x EX. Then the following conditions are equipotent: a) the mapping F is upper semicontinuous at the point x and the set F(x) is compact;
Some properties of topological, metric and Euclidean spaces.
55
b) for every sequence Xk ---+ x of points of the space X every sequence of points Yk E F(Xk), k = 1,2, ... has a subsequence converging to a point of the set F(x). Proof. a~b. This follows from Lemma 5.1 because the point x possesses a countable base of neighborhoods, for instance, the base {O( x, 2-k) : k=1,2, ... }. b ~a. We can consider the stationary sequence Xk == x. When we apply the condition b to this sequence we obtain the compactness of the set F(x). The rest follows from Lemma 5.1. The theorem is proved. • Lemma 5.2. Let cp be a decreasing sequence Fl "2 F2 "2 F3 "2 ... of nonempty subsets of a set A. A generalized sequence IP = {Xa: a E A} of subsets of the space lRn converges with respect to cp to a nonempty convex compact set H ~ lRn in the sense of condition (1.7.1) if and only if H "2 n { cc (H U (u { X a: a E F})): F E cp}. Proof. Necessity. By Lemma 4.3 for every c > 0 the set O£H is convex. For some F E cp H U (U{Xa: a E F}) ~ O£H. By virtue of the convexity of the set O£H we have c(H U (U{Xa: a E F})) ~ O£H. Therefore cc(HU(U{Xa: a E F})) ~ [c(HU(U{Xa: a E F}))] ~ [OcH], n{cc(H U (U{Xa: a E F})): F E cp} ~ [O£H] (for every c > 0). This gives what was required. Sufficiency. Fix an arbitrary point x E H. Assume that the generalized sequence IP does not converge to the set H in the sense of (1. 7.1). By Corollary of Lemma 3.8 there exists a number c > 0 such that for every k = 1,2, ... there exists an index ak E Fk such that X ak \ O£H i 0. Fix a point Xk E X ak \ O£H. The segment connecting the points x and Xk lies in the set cc(H U (U{Xa: a E Fd)). Let tk = sup{s: S E [0,1]' SXk + (1 - s)x E O£H} (recall: the point x belongs and the point Xk does not belong to the set O£H). The point x~ = tkXk + (1 - tk)x belongs to the closure of the set O£H. It does not belong to this set itself. Hence the point lies in the boundary. The boundary of the set O£H is a closed bounded set. By Theoren_ 4.1 it is compact. Therefore the sequence {x~: k = 1,2, ... } has a limit point x* E O£H. We have x' ~ Hand X* E n{ cc(H U (U{ Xc> : a E F})): F E cp} at once. This contradicts our assumptions. Thus our assumption is false and the sequence IP converges to H with respect to cp in the sense of (1.7.1). The lemma is proved. • n Theorem 5.2. Let F : X ---+ lR be a multi-valued mapping of a metric space X into the space lRn , x EX. Let the set F (x) be nonempty compact and convex. Then the following conditions are equipotent: a) the mapping F is upper semicontinuous at the point x; b) F(x) ::2 n{ cc(F(O£x)): c > o}. Proof. The proof consists in referring to Lemma 5.2.
•
CHAPTER 2
56
Lemma 5.3. Let generalized sequences {D",: a E A} and {E", : a E A} of subsets of a normal space X converge in the sense of (1. 7.1) with respect to a directed set
n E",
~
OD
n OE
= (U U V)
n (U U W)
= U U (V
n W)
= UU0 =
u.
This gives what was required. The lemma is proved. • Lemma 5.3 implies immediately: Theorem 5.3. Let multi-valued mappings F, G, with closed values, of a topological space X into a normal space Y be upper semicontinuous (respectively, upper semicontinuous at a point x EX). Then the multi-valued mapping H : X ---+ Y, H(t) = F(t) n G(t), is upper semicontinuous (respectively, upper semicontinuous at the point x). • 6. Connectedness Every topological space has at least two sets which are both simultaneously open and closed. These are the set which coincides with the entire space and the empty set. A topological space X is called connected if it has no proper (i.e., different from X and 0) open-closed (i.e., both open and closed) subsets. In the opposite case, i.e., if (6.1) the space X has a proper open-closed subset, the space X is called disconnected. Denote the proper open-closed subset of the condition (6.1) by F and put G = X \ F. By Theorem 1.3.3 (6.1) implies the conditions (6.2) the space X may be represented as the union of two nonempty disjoint open subsets of it (namely, F and G), and (6.3) the space X may be represented as the union of two nonempty disjoint closed subsets of it
Some properties of topological, metric and Euclidean spaces.
57
(namely, F and G, moreover). By Theorem 1.3.3 condition (6.2) as well as (6.3) implies the fulfilment of condition (6.1). Thus conditions (6.1), (6.2) and (6.3) are equipotent. We have defined the connectedness and the disconnectedness of a topological space. As we usually do with all topological notions we allow one to speak about the connectedness and the disconnectedness of subsets and subspaces of topological spaces too. We then mean the induced topology. Example 6.1. A space consisting of one point is connected. Example 6.2. A segment [a, b], a < b, of the real line is a connected space. Assume the opposite. Then some proper subset M of it is openclosed. Take an arbitrary point t E [a, b] \ M. Denote by s a point of the set M nearest to the point t (see Lemma 3.8). The interval with the endpoints sand t lies in [a, b] \ M. Every neighborhood of the point s intersects the interval. Hence the set M is not open. This contradiction gives what was required. Example 6.3. The subset [0,1] U [2,3] of the real line is disconnected. It is sufficient to refer to (6.3). Lemma 6.1. Let a space X be represented in the form of the union of two disjoint open (respectively, closed) sets F and G. Let a set M ~ X intersect both F and G. Then the set M is disconnected. Proof. The sets MnF and MnG are nonempty and open (respectively, closed) in the subspace M. Their union coincides with M. They are disjoint. Referring to (6.2) (respectively, to (6.3)) completes the proof. • Corollary 1. Let a connected subset M of a topological space be covered by two disjoint open (respectively, closed) subsets. Then the set M lies • entirely in one of them. Corollary 2. Let a connected subset M of a topological space intersect an open-closed set F. Then M ~ F. • Theorem 6.1. Let a topological space X be the union of a family, of its connected subsets and 1= 0. Then the space X is connected. Proof. Fix an arbitrary point x En,. Assume that the space X is disconnected. Take its representation in the form of the union of open sets F and G according to (6.2). For definiteness let x E F. By Corollary 1 of ~ F Lemma 6.1 every element of the family ,lies in F. Therefore X = and G = 0. This contradicts the choice of G. The contradiction shows that Our assumption is false. The theorem is proved. • Example 6.4. A half interval (a, b] of the real line is connected: we apply Theorem 6.1 to the representation (a, b] = U{[c, b]: c E (a, b]}. Likewise a half interval [a, b) is connected. An interval (a, b) is connected: we apply Theorem 6.1 to the representation (a, b) = (a, c] U [c, b), where c is an arbitrary point of the interval (a, b).
n,
U,
58
CHAPTER 2
Theorem 6.2. Let M be a connected subset of a topological space. Then the set [M] is connected. Proof. Assume the opposite. Let [M] = FuG be a representation of the subspace [M] according to (6.2). The set M is dense in the space [M]. The sets F and G are nonempty and open in the space [M]. Thus M n F i= 0 and M n G i= 0. Now Lemma 6.1 implies the disconnectedness of the set M. This contradicts our initial assumption. So our assumption is false. The theorem is proved. • Theorem 6.3. Every connected subset of the real line is either a segment, either a half-interval, or an interval. (A one point subset of the line is regarded as a (degenerate) segment.) Proof. If the set is one point, it is a (degenerate) segment. Let a connected subset M of the real line contain more than one point. Show first that for every two points s < t of the set M the segment [s, t] lies in M. Assume the opposite, i.e., there exists a point c E [s, t] \ M. Then the sets M n (-00, c) and M n (c, (0) are nonempty. Lemma 6.1 implies the disconnectedness of the set M. This contradicts initial assumptions. Thus our additional assumption is false. So the interval (a, b), where a = inf M and b = sup M, lies in set M. The definition of a and b implies that the set M is either the segment [a, b], either the half interval [a, b), either the half interval (a, b], or the interval (a, b). All these sets are connected (see • Examples 6.1, 6.2 and 6.4). The theorem is proved. A connected metric compactum is called a continuum. Theorem 6.3 implies that a subset of the line is a continuum if and only if it is a segment (because in the list of Theorem 6.3 only segments are compact). Singletons are called trivial connected sets. Connected sets containing more than one point are called nontrivial. Theorem 6.4. Let h be an upper semicontinuous (multi-valued) mapping of a connected topological space X into a topological space Y. Let for every point x E X the set h(x) be nonempty and connected. Then the set h(X) is connected. Proof. Assume the opposite. The set h(X) may be represented in the form h(X) = F U G according to (6.2). Corollary 2 of Lemma 6.1 implies the equalities h-1(F) = {x: x E X, h(x) ~ F} and h-l(G) = {x : x E X, h(x) ~ G}. Therefore h-1(F) n h-l(G) = 0. These sets are nonempty. Their union coincides with X. Theorem 1.7.5 implies the openness of these sets. So we have the representation (6.2). Hence X is disconnected. This contradicts the initial assumption. So our additional assumption is false. The theorem is proved. • Corollary 1. The image of a connected topological space under a continuous (single valued) mapping is connected. •
Some properties of topological, metric and Euclidean spaces.
59
Example 6.5. A circle is connected. Consider for simplicity the circle x 2 + y2 = 1. To prove the connectedness it is sufficient to apply Corollary 1 to the mapping p : JR -. JR2, p(t) = (cos t, sin t). Corollary 2. Let h be a continuous mapping of a compactum X onto a connected Hausdorff space Y. Let the preimage of every point y E Y under the mapping h be connected. Then the compactum X is connected. (See Theorem 6.4 and Corollary of Theorem 1.7.8.) • Remark 6.1. Corollary 2 implies the connectedness of the product of two connected compacta. (We consider the projection of the product onto the factor.) However the requirement of the compactness of the factors in fact is superfluous: the product of two connected spaces is connected. To prove that we may consider the multi-valued mapping q being inverse to the projection p of the product onto a factor. In the general case the mapping q need not be upper semicontinuous. However, because the image under the mapping p of every open set is an open set replaces well the assumption of the upper semicontinuity of q. We repeat with corresponding small changes the proof of Theorem 6.4 to obtain what was required. A topological space is called locally connected if every its point possesses arbitrary small connected neighborhood (i.e., for every point x of the space in question and for every neighborhood Ox of this point there exists a connected neighborhood of x lying in Ox). Example 6.6. Not every locally connected space is connected. The space of Example 6.3 is locally connected and disconnected. Example 6.7. Not every connected space is locally connected. In order to construct a corresponding example, denote by lk, k = 1,2, ... , the segment connecting in the plane the points (0,1) and (2- k , 0) and put 10 = {O} x [0,1]. By Theorem 6.1 the set X = U{lk: k = 0,1,2, ... } (Figure 2.3) is connected. The point (0,0) does not possess a connected neighborhood (in the space X) lying in the neighborhood of the radius 1. Let x be an arbitrary point of a topological space X. Denote by ex the union of all connected subsets of the space X conFigure 2.3 taining the point x. The set ex is nonempty because it contains the point x. Theorem 6.1 implies its connectedness. Theorem 6.2 implies its closedness. The set ex is called a connected component (or, shortly, a component) of the point x in the space X. Theorem 6.1 implies that if two component intersect then they coincide. In particular,
60
CHAPTER 2
a point x belongs to the component of a point y if and only if the point y belongs to the component of the point x. Evidently: Theorem 6.5. A connected component of a locally connected space is an open set. • Theorem 6.6. Let, be an open cover of a topological space X consisting of pairwise disjoint nonempty connected sets. Then all connected components of the space X and only them are elements of ,. Proof. By Corollary 2 of Lemma 6.1 every component of the space X lies entirely in an element of the cover ,. It follows from the definition that this component coincides with the corresponding element of ,. By the same arguments every element of, is a component. The theorem is proved. • Theorem 6.7. Let U be an open subset of a locally connected space. Then the set U may be represented as the union of a family , of pairwise disjoint nonempty connected open subset and such a representation is umque. Proof. The subspace U is locally connected and Theorem 6.5 gives the needed representation. Theorem 6.6 implies its uniqueness. The theorem is • proved. Corollary. Let U be an open subset of a locally connected space with countable base. Then the set U may be represented as the union of an (at most) countable family of pairwise disjoint (nonempty) connected open sets and such a representation is unique. (See Corollary of Theorem 1.6.10.) • Example 6.8. Connected components of the Cantor perfect set (see Example 1.6.4) are singletons. A continuous mapping f of (every) segment [a, b] of the real line into a topological space X is called a path in the space X. The point f (a) is called the initial point or first endpoint and the point f (b) is called the end of the or second endpoint of the path f. They say that the path f connects the points f (a) and f (b). Every two nondegenerate segments are homeomorphic (see Example 2.1). Therefore when we have a path f we can take every other segment I of the real line and construct a path with the same endpoints and with the segment I as its domain of definition. The segment being the domain of definition of the path f is symmetric with respect to its middle point. When we use the symmetry in a analogous way we can construct a path having the point f (b) as its initial point and the point f (a) as the end. Lemma 6.2. Assume that there exist a path connecting points a and b and a path connecting points band c in a topological space X. Then there exists a path connecting the points a and c in the space X. Proof. The proof is obvious.
•
Some properties of topological, metric and Euclidean spaces.
61
A topological space is called arcwise connected if any two points of it can be connected by a path. A topological space is called locally arcwise connected if any point of it possesses an arbitrarily small arcwise connected neighborhood. Corollary 1 of Theorem 6.4 and Theorem 6.1 imply that every arcwise connected space is connected (recall: in Example 6.2 we have established the connectedness of a segment). So every locally arcwise connected space is locally connected. Let a space X be locally arcwise connected. For x E X denote by Lx the set of all points of the space X which can be connected to the point x by a path. If a point y belongs to the closure of the set Lx then (every) connected neighborhood Oy of it intersects the set Lx. Lemma 6.2 implies that the set Lx U Oy is arcwise connected. Hence it lies in Lx. In particular, Oy ~ Lx· Thus y E (Lx). In view of the arbitrariness of the point y E [Lx] this means that [Lx] ~ (Lx). Hence Lx = [Lx] = (Lx). So the set Lx is open-closed. By Corollary 1 of Lemma 6.1 the set Lx coincides with the connected component of the point x.
7. Plane regions An open connected subset of the space ]Rn is called a region. By the Corollary of Theorem 6.8 every open subset of the space ]Rn fall into the union of an (at most) countable family of pairwise disjoint regions. This concerns, in particular, the complement ]Rn \ X of an arbitrary compact subset X of the space ]Rn. Let , be the set of connected component of the set ]Rn \ X (we say that X splits the space ]Rn into the sets U E ,). By Lemma 3.7 and Remark 3.2, for some r > 0 the compact urn X lies in the ball 0(0, r). The set ]Rn \ 0, (0, r) ~ ]Rn \ X is (arcwise) connected. By Corollary 2 of Lemma 6.1 it lies in an element W oo(X) of the family,. All other elements of ,lie in the ball 0,(0, r). Hence they are bounded. Let Y be a Hausdorff space, a < b. A path f : [a, b] ~ Y is called simple if it is injective. A path f is called a loop, or a closed path, if f (a) = f (b). The loop (closed path) f is called simple if f (s) =f. f (t) for every pair of different points s, t E (a, b]. If f : [a, b] ~ Y is a simple path, then the mapping f is an embedding (see Theorem 2.1). A continuous mapping of a segment (i.e., a path), of a half interval or of an interval is called a curve. We say that a curve has no self-intersections if it is injective. A closed curve is a synonym for a closed path. A closed curve without self-intersections is a synonym for a simple loop. Let f : [a, b] ~ Y be a loop. The formula
27r(t - a) . 27r(t - a)) h() t = ( cos b -a ' SIn b -a
62
CHAPTER 2
defines a continuous mapping of the segment [a, bj onto the circle S of the center in the origin of coordinates and the radius 1. The continuity of the (single valued) mapping fh- 1 : S --+ f([a, b]) follows from the Corollary of Theorem 1.7.8 (because the composition of two upper semicontinuous mapping is upper semicontinuous). The constructed mapping fh- 1 is also called a closed path, or a loop. If f is a simple loop then the mapping fh- 1 is one to one. In this case by Theorem 2.1 the mapping fh- 1 is an homeomorphism. It is also called a simple closed path, or a simple loop. If f : [a, bj --+ Y is a simple path (respectively, a simple loop) then the set f([a, b]) is called a simple arc (respectively, a simple closed arc). So a simple arc is a subset homeomorphic to a segment. A simple closed arc is a subset homeomorphic to a circle. One of the corresponding implications is proved above; the converse (i.e., if a set is homeomorphic to a circle then it may be covered by a simple loop) is somewhat obvious. For brevity we will say that a simple loop bounds a region V ~ ]R2, if the corresponding simple closed arc is the boundary of the bounded region V. Points f (a) and f (b) are called an initial point and an end, respectively, of the simple arc f([a, b]). If the end of a simple arc coincides with the initial point of another and they do not have any other common points then their union is also a simple arc (with the initial point at the initial point of the first arc and with the end at the end of the second). If the end of a simple arc coincides with the initial point of a second and the end of the second arc coincides with the initial point of the first, and the arcs have no other common points, then their union is a simple closed arc. A simple arc does not split the plane (i.e., its complement is connected or, with the notation of the beginning of this section, the set , consists of the unique element Woo(X), where X denotes the simple arc in question) [Dij.
Theorem 7.1 (Jordan). A simple closed arc splits the plane into two regions and is the boundary of every of them. See [Dij.
•
Notice with reference to Theorem 7.1 that one of two components of the complement to a simple arc X is just the unbounded component Woo(X) and second component is a bounded set.
Theorem 7.2. Let a plane region U be bounded by a simple closed arc. Then there exists an homeomorphism g of the set [Uj onto the disc {(x,y): (x,y) E ]R2, x 2 + y :'( I} such that g(U) = {(x,y): (x,y) E 1l~.2, x 2 + y2 < I}. Proof. The proof of this assertion is based on methods of the theory of functions of a complex variable (Riemann theorem and the theorem on the correspondence of boundaries accompanying the Riemann theorem, see [Maj). •
Some properties of topological, metric and Euclidean spaces.
63
8. Degree of a mapping of a circle into itself The mapping p, p(t) = (cos t, sin t), maps the real line into the circle S ofthe radius 1 and the center in the origin of a rectangular system of coordinates of the plane, (see its graph in Figure 2.4). If we take an arbitrary intervall of the circle S (i.e., an arc without its endpoints) then its pre image under the mapping p falls into the union of pairwise disjoint intervals (0:+ 21fk, (3+ 21fk), where 0 < (3 -0: :( 21f and k = 0, ±1, ±2, ... , moreover the restriction of the mapping p to every of these intervals is an homeomorphism onto the intervall (Figure 2.4). We will consider the set e(X, JR.) (respectively, e(X, S)) of all upper semicontinuous mapping of a compactum X in the line JR. (respectively, in the circle S) with proper connected compact subset as values. By Theorem 6.3 values of mappings from e(X, JR.) are segments of the real line. The same theorem implies that values of mappings from e(X, S) are segments (i.e., arcs with their endpoints) of the circle S. A mapping 9 E e(X, JR.) is called an e-lifting of the mapping I E e(X, S), if 1= pg.
Figure 2.4
Let -00 < e < d < 00. When we consider mappings of a eompactum X, mappings from e(X x [e, d], JR.) and e(X x [e, d], S) are called e-homotopies. Mappings 11,12 E e(X, JR.) (respectively, f1, h E e(X, S)) are called e-homotopie if there exists an e-homotopy F E e(X x [e, d], JR.) (respectively, F E e(X x [e, d], S)) such that II (x) = F(x, e) and h(x) = F(x, d) for every x E X. All nondegenerate segments of the line are homeomorphic. So in this definition the segment [e, d] may be replaced by every other nondegenerate segment of the line, for instance, by the segment [0, 1]. Likewise in this definition the order of 11 and h is not essential. Every mapping 1 E e(X, IR) U e(X, S) is e-homotopic to itself (a corresponding e-homotopy may be defined by the formula F(x, t) = l(x)). If 11 is e-homotopic to 12 and is 12 e-homotopic to h, then 11 is e-homotopic 13 (a corresponding
64
CHAPTER 2
homotopy may be defined by the formula for x E X, t E [0,1], for x E X, t E [1,2]' where Fl is an e-homotopy from the mapping fl to the mapping hand F2 is an e-homotopy from f2 to h, see Theorem 1.7.6). Lemma 8.1. Let FE e(X x [c,d],S) and 9 E e(X,IR). Let F(x,c) = pg(x) for every point x E X. Then there exists an unique e-lifting G E e(X x [c, d], IR) of the e-homotopy F satisfying the condition
G(x, c) = g(x) for every point x E
x.
Proof. 1. Let us prove the uniqueness of the e-homotopy G. Let two e-homotopies G 1 and G 2 satisfy the condition imposed on the homotopy G and (x, s) E X x [c, d]. For every point t E [c,d] we have p(G 1 (x,t)) = F(x,t) = p(G 2 (x,t)). The lengths of the segments G 1 (x, t) and G 2 (x, t) are less than 271". Thus G 1 (x, t) = G 2 (x, t) + 271"k(t), where k(t) = 0, ±1, ±2, .... The upper semicontinuity of the mappings G 1 and G 2 implies the continuity of the function k. Values of the function k are integers. The set of integers does not contain nontrivial connected subsets. We obtain from the continuity of the function k and Corollary 1 of Theorem 6.4 that the function k is constant. Since k(c) = 0, k == 0. Hence G 1 (x, t) = G 2 (x, t). This gives what was required. II. Let us prove the existence of the e-homotopy G under the strong additional restriction F(X x [c, d]) i- S. Fix arbitrary points 0 E S\F(X x [c,d]) and a E p-l(O). The preimage of the set S\ {O} under the mapping p falls into the union of pairwise disjoint intervals I j = (a + 271"j, a + 271"(j + 1)), j = 0, ±1, ±2, .... The restriction of the mapping p to every such interval is an homeomorphism onto the set S \ {O}. Let qj : S \ {O} -Y I j denote the inverse homeomorphism. Since g(X) ~ IR\p-l(O), Corollary 1 of Lemma 6.1 and the membership 9 E e(X, IR) imply the representation of the compactum X as the union of its pairwise disjoint open subsets Xj = g-l(1j), j = 0, ±1, ±2, .... Since each of them is equal to the complement of the union of all the others, these sets are closed. Let G(x, t) = qjF(x, t) for x E Xj, where j = 0, ±1, ±2, .... By virtue of our remarks the mapping G satisfies all required conditions. III. Make weaker the restriction imposed in II and pass to the condition F(X x {t})
i- S
for every t E [c, d].
Some properties of topological, metric and Euclidean spaces.
65
For every t E [e, d] fix a point O(t) E S \ F(X x {t}). The set F-I(O(t)) does not intersect the set X x {t}. By Theorem 1.7.5 it is closed in the product X x [e, d]. By virtue of Theorem 1.4 for some t5(t) > 0 the set F-I(O(t)) lies outside the set X x 06( tl t. Hence F(X x 06(t)t) <;;; S\ {O(t)}. Let M denote the set of all m E [e, d] such that: for the e-homotopy F IXx[c,m] there exists an e-lifting G m satisfying all the above conditions but with the change of the segment [e, d] to the segment [e, m] <;;; [e, d]. The set M is nonempty because it contains the point e. The definition of the set M immediately implies that it has the form of a half interval [e, 'Y) or of a segment [e, 'Yl, where 'Y E [e, d]: if m E M and e ~ ml ~ m then ml E M, because we can take as G m, the restriction of the e-homotopy G m to the set X x [e, ml]' By I the mappings G m are defined uniquely. Take an arbitrary point tl of a (nonempty) set M n 06h)'Y' Let t2 be an arbitrary point of the set ['Y, d] n06h)'Y. We have F(X x [tl' t 2]) <;;; S\ {O("()}. The result of II is applicable to the e-homotopy F* = F IXX[t"t2] as F, to the mapping g* : X - t JR, g*(x) = G tl (x, h) as 9 and to the segment [tl, t 2 ] instead of [e, d]. Let G* be a e-lifting existing by II. Define the mapping
G(x t) = {GtG*(x,(x,t)t) l
t2'
t E [e, tIl, X, t E [tl, t 2l,
if x EX, if x E
on the set X x [e, t2]' The upper semieontinuity of the mapping G t2 follows from Theorem 1.7.6. Thus t2 E M. In view of the arbitrariness of the point t2 E [,,(, d] n 06h)'Y this implies the inclusion 06h)'Y <;;; M. The definition of'Y now implies that 'Y = d and M = [e, d]. So we have proved the existence of an e-lifting under the additional assumption. IV. Pass to the general case. Let x EX. By virtue of the upper semieontinuity of the mapping F (and Theorem 1.7.3) for an arbitrary t E [e, d] there exists a Tychonoff neighborhood 0; x Ot of the point (x, t) such that F( 0; x Ox) =f S. The open cover {Ot: t E [e, d]} of the segment [e, d] contains a finite sub cover {Otl , . .. ,Otd. Put Ox = n{ 0;1, ... ,O;k}. Every point t E [e, d] belongs at least to one of the sets Ot i . Therefore
(8.1) Apply Lemma 1.6.1 to the point x and to the set X \ Ox. Lemma 1.6.1 implies the existence of a neighborhood V x of the point x such that
66
[Vx]
CHAPTER 2 ~
Ox. By virtue of(8.1) the mappings FX
F I[Vxjx[c,dj andg = gl[V,j satisfy the condition imposed in II on F and g, respectively (with [Vx] as X). Let GX : [V x] X [c, d] - t ~ be an e-lifting of the mapping FX existing by II. If y E [Vxd n [VX2] and i E [c, d], then by I GXl(y, i) = G X2(y, i). Hence the formula G(y,i) = GX(y,i) for y E [Vx] uniquely defines the mapping G : X x [c, d] - t ~. Theorem 1.7.7 implies its upper semicontinuity. The fulfilment of all other conditions on G is obvious. The lemma is proved . • When the compactum X consists of one point we obtain: Corollary. Lei f E e([c, d]' S), I be a segment of the real line, p(1) = f(c). Then there exists a unique e-lifting h of the mapping f satisfying the condition h(c) = I. • Let us now define the notion which is mentioned in the title of this section. For every mapping g E e(S, S) we can apply Corollary of Lemma 8.1 to the mapping f = g . p I[O,2-11} [0,21f] - t S (we can take as I an arbitrary segment satisfying the condition p(I) = f(c)). If h is an e-lifting of the mapping f existing by Corollary of Lemma 8.1 then ph(O) = f(O) = f(21f) = ph(21f). Therefore the segment h(21f) has the form 1+ 21fk, where k = 0, ±1, ±2, .... The number k is called the degree of the mapping g. In the definition of the degree of a mapping we had an arbitrariness in the choice of the e-lifting h (or, to be precise, an arbitrariness in the choice of the segment 1). Every other e-lifting hl of the mapping f has the form hl (t) = h(i) + 21fi, i = 0, ±1, ±2, ... (this follows from the uniqueness of the e-lifting under restrictions of Corollary). Therefore hi (21f)
=
X
= h(21f) + 21fi = 1+ 21fk + 21fi = hl (21f) + 21fk,
i.e., the value of the degree of a mapping does not depend on the choice of its e-lifting. Every single valued continuous mapping 9 : S - t S belongs to e(S, S). Its degree corresponds to the number of entire rotations of the point g(i), when i runs the circle one time. The direction of the rotations defines the sign of the degree. Theorem 8.1. Mappings
h3 (c) = G(21f, c) = g(21f) = g(O)
+ 21fk = G(O, c) + 21fk = hl (c) + 21fk = h 2 ( c).
Some properties of topological, metric and Euclidean spaces.
67
The mappings h2 and h3 are e-liftings of the mapping a(t) = F(O, t). The uniqueness of such an e-lifting (see Corollary of Lemma 8.1) guarantees the coincidence of the mappings h2 and h 3. Thus
g*(27T) = G(27T, d) h3(d) = h2(d) = hl(d) + 27Tk = G(O, d) + 27Tk = g*(O) + 27Tk. =
Hence the degree of the mapping 1f2 is also equal to k. Sufficiency. Let k be the degree of the mappings Ifl and 1f2. Let 'l/Ji = lfiP l[o,27rr [0,27T] -+ 8 for i = 1,2. Let hI, h2 be e-liftings for Ifl and 1f2, respectively. Evidently the mapping H (s, t) = hI (s ) (1 - t) + h2 (s )t belongs to e([O, 27T] x [0,1]' JR) (the upper semicontinuity of the mapping H may be easily established, for instance, with the help of Theorem 5.1). The mapping F =pH belongs to e([0,27T] x [0,1],8). For every t E [0,1]
=
+ h2(27T)t) p«hl(O) + 27Tk)(1 - t) + (h 2(0) + 27Tk)t) p(h l (O)(1 - t) + h 2(0)t) + 27Tk) = p(h l (O)(1
=
F(O, t).
F(27T, t) = p(h l (27T)(1 - t) =
- t)
+ h2(0)t)
Therefore there exists an unique mapping G : 8 x [0,1] -+ 8 satisfying the condition F(x, t) = G(p(x), t) for every x E [0,27T] and t E [0,1]. Evidently G E e(8 x [0,1]' 8) (the upper semicontinuity of the mapping G follows easily from the upper semicontinuity of the mapping 1-1 : 8 x [0,1] -+ [0,27T] X [0,1], where 1 : [0,27T] X [0,1] -+ 8 x [0,1]' l(x, t) = (p(x), t), see Corollary of Theorem 1.7.8). It remains to notice that the e-homotopy G connects the mappings Ifl and 1f2. The theorem is proved. • A continuous (single valued) mapping F : X x [c, d] -+ Y of the product of a topological space X and a segment [c, d] of the real line into a topological space Y is called a homotopy connecting the mappings lc(x) = F(x, c) and Id(x) = F(x, d). Mappings are called homotopic if there exists a homotopy connecting them. Evidently a single valued e-homotopy is an homotopy. In the case of the single valuedness of the mappings Ifl and 1f2 in Theorem 8.1 the mapping G in the second part of its proof is single valued too. Therefore the last remark and Theorem 8.1 imply: Theorem 8.2. Continuous (single valued) mappings 1f1, 1f2 : 8 -+ 8 are homotopic (i.e., may be connected by a homotopy) if and only if their degrees coincide. • The formula x* (t) = x cos t - y sin t { y* (t) = x sin t + Y cos t
68
CHAPTER 2
defines a continuous mapping S x IR ----) S. For a fixed t E IR we have a rotation of the circle S about the angle t. Therefore the identity mapping S ----) S and the rotation of the circle S about an arbitrary angle are homotopic. This fact and Theorem 8.1 imply that in the definition of the degree of a mapping the segment [0,21f] may be changed to every other segment of the line of the length 21f. Mappings which are homotopic to constant mappings are called contractible. The degree of a constant mapping (of the circle into the circle) is equal to zero. Thus: Example 8.1. A continuous mapping of the circle into the circle is contractible if and only if its degree is equal to zero. Example 8.2. Identify the circle S with the circle {z: z E C, Izl = 1} in the complex plane C. It is not difficult to check that the degree of the mapping cp : S ----) S, cp(z) = zn, is equal to n. (Recall the Euler formula (cos cp + i sin cp) n = (cos ncp + i sin ncp )). Example 8.3. 'Fundamental Theorem of Algebra'. Every polynomial I(z) = zn+an_lzn-l + .. ·+alz+aO (n = 1,2, ... ) with complex coefficients has a complex radical. Let A = nmax{lan_ll, ... ,lall,laol,l}. Since An> lan_lIAn-l + ... + lallA + ao, the (continuous) function F(x, t) = xn + t(an_lx n- l + ... + alx + ao) does vanish for Ixl = A and 0 :::; t :::; 1. Thus the formula
(z, t) = F(Az, t)(IF(Az, t)l)-l, where Izl = 1 and 0 :::; t :::; 1, defines an homotopy. It connects the functions cp(z) = zn and 'IjJ(z) = I(Az)(I/(Az)I)-l. We noticed in Example 8.2 that the degree of cp is equal to n. By Theorem 8.2 the degree of'IjJ is also equal to n. Assume now that the polynomial 1 has no radicals. Then the formula *(z, t) = f(tz)(lf(tz)l)-l, where Izl = 1 and 0 :::; t :::; A, defines an homotopy connecting the mapping 'IjJ and the constant mapping 'IjJ* == 1(0)(1/(0)1)-1. We noticed in Example 8.1 that the degree of a constant mapping is equal to zero. Therefore the result of the previous paragraph contradicts Theorem 8.2. Thus our assumption is false. This gives what was required. Lemma 8.2. Let mappings I, g E e(X, S) of a compactum X into the circle S satisfy the condition
I(x)
~
g(x) for every point x E X.
Then the mappings f and g are e-homotopic. Proof. We can define an e-homotopy connecting the mappings g by the formula
f(X) h(x, t) = { g(x)
for x E X,D:::; t < 1, for x EX, t = 1.
f and
Some properties of topological, metric and Euclidean spaces.
69
That it belongs to e(X x [0,1]' S) is obvious. • Corollary. Let mappings fI' h E e(X, S) of a compactum X into the circle S satisfy the condition:
for every x E X the set fl (x) U opposite points.
h (x)
does not contain diametrically
Then the mappings fl and f2 are e-homotopic. To obtain the corollary let us notice that under our assumptions for every x E X the set S \ (fl (x) U h (x)) consists of one or two arcs. But only one of them has a length greater than the length of half the circle. Denote by g(x) the complement to this arc. Evidently g E e(X, S). By Lemma 8.2 the mapping g is e-homotopic both to fl and f2' This gives what was required. • The next remark concerns the fulfilment of the condition: (8.2) every (single valued) continuous mapping f point (i.e., a point x E X such that f(x) = x).
:X
--*
X possesses a fixed
If a space X satisfies (8.2) and a space Y is homeomorphic to the space X then the space Y satisfies the condition (8.2) too. Theorem 8.3. Every continuous mapping of a closed disc into itself
possesses a fixed point. Proof. By virtue of the previous remark it is sufficient to consider the disc D = {( x, y) : (x, y) E lie, x 2 + y2 ~ 1} of the radius 1 and the center in the origin of the rectangular system of coordinates. Assume the opposite. Let a continuous mapping f : D --* D have no fixed points. For zED put u(z) = z - f(z) and v(z) = u(z) (1Iu(z)II)-I. Since (z, f(z)) ~ Izl . If(z)1 ~ 1,
(u(z), z) = (z - f(z), z) = (z, z) - (f(z), z)
~
0
for Ilzll = 1. Thus the points z and v(z) are not diametrically opposite. The degree of the identity mapping S --* S is equal to 1. By the Corollary of Lemma 8.2 and Theorem 8.1 the degree of the mapping 9 = v Is is equal to 1 too. The homotopy G(z, t) = v(tz), where z E S, 0 ~ t ~ 1, connects the mapping 9 and the constant mapping h == v(O). By Theorem 8.2 the degree of the mapping 9 is equal to zero. This contradicts the previous calculation. Thus our assumption is false. The theorem is proved. • The following more general assertions are true too.
70
CHAPTER 2
Theorem 8.4 (Brouwer Fixed Point Theorem). Every continuous mapping of a closed ball of the Euclidean space (of arbitrary dimension) into itself possesses a fixed point. See [HW]. • Theorem 8.5 (Schauder Fixed Point Theorem). Every continuous mapping of a compact convex subset of a Banach space (see below §3.1) into itself possesses a fixed point. See [Ed]. • A mapping j* E e([c, d], S) satisfying the condition j*(c) = j*(d), is called an e-loop. By the Corollary of Lemma 8.1 the e-loop j* possesses an e-lifting h*, and as ph*(e) = j*(e) = j*(d) = ph'(d) then hOed) = h*(e) + 27fk for some k = 0, ±1, ±2, .... The number k is called the degree of the e-loop j*. As for the degree of a mapping of the circle the value of the degree of the e-loop j* does not depend on the choice of the e-lifting h'. With the notation of the definition of the degree of a mapping of the circle the degree of the mapping g is equal to the degree of the e-loop f. On the other hand, if a mapping q : [e, d] -+ S is defined by the formula _ ( 27f(t - c) . 27f(t - C») q (t ) - cos d ' sm d -c -e
then the degree ofthe e-loop j* is equal to the degree of the mapping j*q-l. The proof reduces to a simple comparison of definitions. The use of the notion of a e-loop simplify a little the calculation of the degree of mappings. It is based on the obvious Lemma 8.3. Let the degrees of e-loops fl E e([b, c], S), h E e([e, d], S) be equal, respectively, to kl and k 2. Let fl (c) = h (c). Then the degree of the e-loop f E e( [b, d]' S), for t E [b, e], for t E [e, d], is equal to kl + k 2 . Proof. Let hI E e([b, c], JR) be an e-lifting for fl' Fix an e-lifting h2 of e-loop h according to Corollary of Lemma 8.1 in order to satisfy the condition h2 (c) = hI (c). It remains to notice that the mapping h(t) = {hI (t) h2(t)
for t E [b, c], for t E [e, d],
is an e-lifting of the e-Ioop f and h(d) = h2(d) = h2(e) + 27fk2 = hl(e) + 27fk2 = hI(b) + 27fkl + 27fk2 = h(b) + 27f(kl + k2)' The lemma is proved .•
CHAPTER 3
SPACES OF MAPPINGS AND SPACES OF COMPACT SUBSETS
Usually in accounts of the theory of Ordinary Differential Equations in the description of topological properties of sets of solutions we use the notion of the uniform convergence of sequences of functions. We will discuss related topological structure in §§ 1 and 2. In fact its correspondence with this circle of questions is not very good because it deals with sets of functions with a fixed domain. In the definition of a solution of a differential equation we do not fix domains of definition. Functions with different domains are called solutions of the same equation. In order to describe topological properties of such sets of functions more completely, but rather briefly, we will introduce in §5 the space of partial mappings. In first few sections we consider some fundamental notions which are helpful in study of properties of the new space. See also [Ku1].
1. Metric and norm of uniform convergence Let X be an arbitrary set. Let (Y, p) be a metric space. We considered in Example 1.4.4 the space B(X, Y) of all bounded mappings of the set X into the space Y. The metric d of Example 1.4.4 is called the metric of uniform convergence. Lemma 1.1 Let (with the above notation) the space Y be complete. Then the space B(X, Y) is complete too. Proof. Let {Jk: k = 1,2, ... } be a fundamental sequence of elements of the space B(X, Y). By virtue of the estimate P(fi(X), fj(x)) ~ d(j;, fj) the sequence {Jk(X): k = 1,2, ... } is fundamental for every element x E X. By virtue of the completeness of the space Y the sequence converges to a point f (x). The verification of the facts of the function f just defined is bounded (and hence belongs to the space B(X, Y)) and that d(ik, J) ----+ 0 will complete the proof of the lemma. We leave it to the reader as a simple exercise. • Assume now in addition that X is a topological space. Denote the set of all bounded continuous mappings of the space X into the space Y by BG(X, Y).
72
CHAPTER 3
Lemma 1.2. Under the conditions mentioned the set BC(X, Y) is closed in the space B(X, Y). Proof. The proof of this assertion reduces to the property that the limit of an uniformly convergent sequence of continuous functions is a continuous • function, see [AI]. Lemma 1.3. A closed subspace of a complete metric space is complete (with respect to the induced metric). Proof. The proof is obvious. Lemmas 1.1-3 immediately imply: Theorem 1.1. The space BC(X, Y) of all continuous bounded mappings of a topological space X into a complete metric space Y is complete (with respect to the metric of uniform convergence). • Denote by C(X, Y) the set of all continuous mappings of a topological space X into a topological space Y. (So, BC(X, Y) = B(X, Y)
n C(X, Y).)
If now a space X is compact and Y is a metric space then by Theorem 1.7.8 and Lemma 2.3.7 C(X, Y) = BC(X, Y). Thus in this case the metric of uniform convergence is defined on the entire set C(X, Y) and Theorem 1.1 implies Corollary. The space C(X, Y) of all continuous mappings of a compact space X into a complete metric space Y is complete (with respect to the metric of the uniform convergence). • Often we consider the space of mappings into a normed space. This situation is already met when we consider the space of real functions. In the corresponding assumptions the mapping space may be considered as a normed space, moreover its norm generates the metric of the uniform convergence. Let X be an arbitrary set and Y be a normed space. For an arbitrary mapping f : X --t Y let us put Ilfll = sup{lIf(x)lI: x E X}. A norm of an element of the space Y is nothing else than the distance from this element to the zero of the space Y. By virtue of the remarks in §2.3 the condition of the boundedness of a mapping f : X --t Y may be written as IIfll < 00. We can introduce in a natural way a linear structure on the space of mappings from X into Y. For two mappings f, g : X --t Y their sum is defined by the formula (J + g)(x) = f(x) + g(x). If a is an arbitrary number the mapping af is defined by the formula (aJ)(x) = a . f(x). With the notation described the function that associates the number IIfll to a mapping f E BC(X, Y) is a norm. The verification of this fact is not difficult. Thus the space BC(X, Y) turns out to be normed. A complete (in the sense of the metric related with the norm) normed space is called a Banach space. Theorem 1.1 implies:
Spaces of mappings and spaces of compact subsets.
73
Theorem 1.2. The space BC(X, Y) of all continuous bounded mappings of a topological space X into a Banach space Y is Banach. • As in the case of Theorem 1.1 we obtain Corollary. The space C(X, Y) of all continuous mappings of a compact space X into a Banach space Y is a Banach space. • We can introduce in a natural way the notion of a series Ul +U2 +U3 + ... and of their convergence as the existence of the limit of the sequence of partial sums Sk = UI + ... + Uk. The limit is called the sum of the series under consideration. Theorem 1.3. Let UI + U2 + U3 + . .. be series of elements of a Banach space Y. Let the numerical series IluI II + IIu2 II + IIu3 II + . .. converge. Then the series UI + U2 + U3 + ... converge too. Proof. By virtue of completeness of the space Y it is sufficient to show that the sequence of partial sums of the series UI + U2 + U3 + . .. is fundamental. It follows easily from the inequality
and from the convergence of the series IIuI II + II u 2 II + II u 3 II + ... (the convergence of the series implies that the sequence of partial sums is fundamental, see §1.6). The theorem is proved. • By virtue of Corollary of Theorem 1.2, Theorem 1.3 may be applied to series of continuous real functions on a compact space. We speak in this case about uniform convergence and about uniformly convergent series. 2. Compact open topology Consider now the possibility of introducing a topology on the set C(X, Y) in the general case, i.e., without the assumption of the previous section about the presence of a norm or a metric on the space Y. We can introduce a topology on the set C(X, Y) in different but quite natural ways. We will consider only one, but it corresponds the most to the circle of questions under consideration. Let X and Y be topological spaces. For K ~ X and U ~ Y denote O(K, U) = {f: f E C(X, Y), f(K) ~ U}. Denote by f3 the family of all sets of the form O(K, U), where K runs over the set of all closed compact subsets of the space X and U runs over the set of all open subsets of the space Y. The family f3 is a sub-base of a topology on the set C(X, Y). It is called the compact open topology on C(X, Y). Notice that our condition defines the topology uniquely. Theorem 2.1. Let X be a compactum and (Y, p) be a metric space. Then the compact open topology on the set C(X, Y) is generated by the metric of the uniform convergence.
74
CHAPTER 3
Proof. Denote by T the compact open topology on the set C(X, Y). Denote by d the metric of uniform convergence on the set C(X, Y). We can state our problem as follows: prove that the identity mapping h : (C(X, Y), d)
--t
(C(X, Y), T)
is an homeomorphism. We will do this with the help of Theorem 1.7.3. I. Let f E O(K, U), where K is a (nonempty) closed compact subset of the space X and U is a (nonempty) open subset ofthe space Y. By Theorem 1.7.8 the set f(K) is compact. By Corollary of Lemma 2.3.8 02cf(K) ~ U for some E > O. Let now g E O(f, E) (notice that in the first case the 2Eneighborhood of the set is taken in the space Y and in the second case the E-neighborhood of the point is taken in the space C(X, Y)). For every point t of the set K and for every point y of the set Y \ U
p(g(t), y)
~
p(f(t), y) - p(f(t), g(t))
~ 2E - E
= E.
In view of the arbitrariness of the point y E Y \ U this implies that E U (Figure 3.1). The point t E K is also chosen arbitrarily. Therefore g E O(K, U). We have established the continuity of the mapping h.
g(t)
Figure 3.1
II. Let f E C(X, Y) and E > O. Let Ox = !-I(Og;'3!(X)) for x E X. The open (see Theorem 1.7.5) cover {Ox: x E X} of the compact space X has a finite subcover {OXI,"" Oxd. By Lemma 2.3.3 and Theorem 1.7.5 the set Ki = !-I([Og/3!(Xi)]), i = 1, ... , k, is closed. Theorem 1.6.2 implies its compactness. Evidently Ki ;:::> OXi' Therefore the family {K I , ... ,Kd covers the space X.
Spaces of mappings and spaces of compact subsets.
75
Let U = n{O(Ki' 02f/3f(Xi)) : i = 1, ... , k}. Since f(Ki) = [O£/J!(Xi))] <;:;; 02£/3f(Xi), fEU. Take arbitrary 9 E U and x EX. For some i = 1, ... ,k we have x E K i ,
p(f(x),g(x)) ~ p(f(x), f(Xi))
E
2E
+ P(f(Xi),g(X)) < 3 +:3
= E
(Figure 3.2). This implies that the function 9 lies in an E-neighborhood of the function f (with respect to d). So we have established the continuity of the mapping h- 1 • The theorem is proved. •
Xi
Figure 3.2
Theorem 2.2. Let a locally compact space X have a countable base. Let Y be a metric space. Then the space C(X, Y) (with the compact open topology) is metrizable. Proof. Let f3 be (at most) a countable base of the space X. The set f30 of all elements of the base f3 with compact closures is (at most) countable: f30 = {Ui : i = 1,2, ... }. The local compactness of the space X implies that f30 is a base of X. For every k = 1,2, ... the set Mk = [U{Ui : i = 1, ... ,k}] is compact. By Theorem 2.1 the space C(Mk' Y) is metrizable. By Theorem 2.2.2 the space Z = IT {C(MkJ Y): k = 1,2, ... } is metrizable. The definition of the compact open topology implies the continuity of the mapping Pk : C(X, Y) - t C(Mk' Y), Pk(Z) = ZIMk ' k = 1,2, .... By Theorem 2.1.1 the diagonal product P : C(X, Y) - t Z of the mappings Pk, k = 1,2, ... , is continuous. The mappings P is injective. The theorem will be proved when we verify the continuity of the inverse mapping Q : P(C(X, Y)) - t C(X, Y). An arbitrary element of the above sub-base of the space C(X, Y) has the form O(K, U), where K is a closed compact subset of the space X and U is
CHAPTER 3
76
an open subset of the space Y. By virtue of Theorem 1.7.3 it is sufficient to prove the openness of the set Q-l (O(K, U)). The compactness of the set K and the equality UfJo = X imply that K ~ Mk for some k = 1,2, .... The set W = {h: h E C(Mk' Y), h(K) ~ U} is open in the space C(Mk' Y) by the definition of the compact open topology. We have
Q-'(O(K, U))
~
(It
C(M" Y) x W x
ilL i,
C(M V)) n P(C(X, V)).
This gives what was required. The theorem is proved. • The construction in the proof of the theorem allows us to characterize the compact open topology as the topology of the uniform convergence on compact subsets of the space X. In particular: Theorem 2.3. Under the hypotheses of Theorem 2.2 a sequence {Ii : i = 1,2, ... } ~ C(X, Y) converges (with respect to the compact open topology) to a function f E C(X, Y) if and only if for every closed compact subset K of the space X the sequence {Ii IK: i = 1,2, ... } converges uniformly to the function f IK. Proof. Necessity follows from Lemma 1.7.2 when we apply it to the (obviously continuous single valued) mapping associating to a function from C(X, Y) its restriction to the set K. Sufficiency follows from Lemma 2.1.2 when we apply it to the embedding P : C(X, Y) - t Z from the proof of Theorem 2.2. The theorem is proved. • Theorem 2.4. Let «I> : T x X - t Y be a continuous mapping of the product of topological spaces T and X into a topological space Y. Let 'Pt(x) = «I>(t, x) for t E T, x E X. Then the mapping F : T - t C(X, Y), F(t) = 'Pt, is continuous. Proof. Let to E T and F(t o) E O(K, U), where K is a closed compact subset of the space X and U is an open subset of the space Y. We have «I> ( {to} x K) = 'Pta (K) ~ U. By virtue of the continuity of the mapping «I> we can associate to every point x of the set K its neighborhood Ox and a neighborhood V(x) of the point to in a manner that «I>(V(x) x Ox) ~ U. The open cover {Ox: x E K} of the compact set K has a finite subcover {OXl, ... , Ox;}. Put V = V(xd n··· n V(Xi). Let t E V and x E K. For some j = 1, ... ,i the point x belongs to OXj. Therefore
In view of the arbitrariness of the point x E K this implies the inclusion 'Pt(K) ~ U. It means that F(t) = 'Pt E O(K, U). Referring to Theorem 1. 7.3 completes the proof. •
Spaces of mappings and spaces of compact subsets.
77
3. Vietoris topology Let X be a topological space. Denote by exp X the set of all nonempty closed subsets of the space X. For U I , . .. ,Uk s:;; X put O(UI , ... ,Uk) = {F: FE expX, F s:;; Uf=IUi , F nUl i- 0, ... ,F n Uk i- 0}. Notice two particular cases. For U s:;; X O(U) = {F: F E expX, F s:;; U} and O(U, X) = {F: F E expX, F n U i- 0}. The following equality is obvious: (3.1) On the set exp X introduce a so called Vietoris topology. It is the smallest topology, in which all set of the forms O(U) and O(U, X) are open (i.e., sets of the form mentioned constitute a sub-base of the Vietoris topology). Here U runs over the topology of the space X. Without any detailed discussion let us notice that the openness of sets of the form O(U) corresponds to the notion of convergence in the sense of condition (1. 7.1), and the openness of sets of the form O(U, X) corresponds to the notion of convergence in the sense of the lower limit. By virtue of (3.1) sets of the form O(UI , · · · , Uk), are open in the Vietoris topology. They are called Vietoris neighborhoods (of their points). In fact, Vietoris neighborhoods constitute a base of the Vietoris topology because elements of the above sub-base (the sets O(U) and O(U, X)) are Vietoris neighborhoods too, and, on the other hand, if FE O(UI , . . . , Uk) n O(VI , ... , Vi) then FE O(WI ,· .. , W m ) s:;; O(UI , .. ·, Uk) n O(VI , . . .
,
Vi),
where {WI,"" W m } is the renumbered family {Ui n Vi: i = 1, ... , k, j = 1, ... ,I, F n Ui n Vi i- 0}. The set exp X with the Vietoris topology is called the space of (nonempty) closed subsets of the space X. We will not give an account here of the large ammount of theory related to the notion introduced, see [Ku1], [FF], [En]. We will restrict ourselves to some technical remarks. Let 9 : X -> Y be a continuous mapping of a topological space X into a topological space Y. It is naturally to define the mapping g* of the space exp X as its domain by putting g* (F) = g(F). However in this way we do not always obtain a mapping of the space exp X into the space exp Y because the image of a closed set need not to be closed. In order to pass over this case consider the set E(g*) of all closed subsets of the space X, images under the mapping 9 of which are closed in Y. Consider the set E(g*) with the topology induced by the Vietoris topology. Although the space E(g*) in the general case need not coincide with the space expX, often it turns out to be quite large. So, if the spaces X and Yare Hausdorff, by virtue of
CHAPTER 3
78
Theorems 1.6.1 and 1.7.8 the space E(g*) contains all (nonempty) compact subsets of the space X. Theorem 3.1. Let 9 be a continuous mapping of a topological space X into a topological space Y. Then the mapping g* : E(g*) --+ exp Y is continuous.
Proof. The proof reduces to a direct verification. Let F E E(g*). Let O(UI , ... , Uk) be an arbitrary Vietoris neighborhood of the point g*(F). For VI = g-I(UI ), ... , V k = g-I(Uk ) we have
Therefore F E o (v;. , ... , Vk). This gives what was required. The theorem is proved. • Theorem 3.2. Let X be a regular space. Let generalized sequences 8 = {D,,: a E A} ~ expX and c = {E,,: a E A} ~ expX converge (in the Vietoris topology) with respect to a directed set cp to D and E, respectively. Let D" ~ E" for every a E A. Then D ~ E. Proof. Take an arbitrary t E X \ E. The regularity of the space X implies the existence of disjoint neighborhoods Ot of the point t and OE of the set E in the space X. By virtue of the convergence of the sequence c to E with respect to cp there exists F E cp such that E" E O(OE) for every a E F. Hence D" ~ E" ~ OE, D" nOt = 0 and D" rf. O(Ot, X). By virtue of the convergence of the sequence 8 to D with respect to cp the open set O(Ot, X) cannot be a neighborhood of the point D, i.e., D n Ot = 0 and t rf. D. This gives what was required. The theorem is proved. • Theorem 3.3. Let X be a normal space. Then the set expc X of all connected closed subsets of the space X is closed in the space exp X. Proof. By Theorem 1.3.3 it is sufficient to prove that the set H = exp X \ expc X of all disconnected closed subset of X is open in the space exp X. Let Fo E H. The set Fo may be represented as the union of two nonempty closed disjoint subsets FI and F 2 . By virtue of the normality of the space X there are disjoint neighborhoods UI and U2 of the sets FI and F 2, respectively. The set O(UI , U2) is a neighborhood of the point Fo in the space exp X. Every F E 0 (U I , U2 ) is equal to the union of two nonempty disjoint closed subsets F \ U2 = F nUl and F \ U I = F n U2· Therefore FE H, O(UI , U2 ) ~ H. The theorem is proved. • Corollary. Let FI ;;2 F2 ;;2 F3 ;;2 ... be a non-increasing sequence of nonempty connected compact subsets of a metric space X. Then the set F = n{Fk: k = 1,2, ... } is connected.
To obtain the corollary it is sufficient to restrict ourselves to the case FI = X. We have F = lim tOPk--+oo Fk (see Example 1.5.3). By Corollary of Theorem 1.6.3 the set F is nonempty. Let U be an open subset of the space X. If F E O(U, X), then the definition of the lower limit implies
Spaces of mappings and spaces of compact subsets.
79
that Fk E O(U, X), beginning with some k. If F E O(U), then by Theorem 1.7.2 Fk E O(U), beginning with some k. By Lemma 1.7.3 Fk --+ F in the Vietoris topology. This allows us to obtain the corollary. •
4. Hausdorff metric Let X be a Hausdorff space. By Theorem 1.6.1 every compact subset of the space X is closed, i.e., the set expc X of all nonempty compact subsets of the space X lies in exp X. The Vietoris topology was defined first on the set exp X. Now it induces a topology on the set expc X. The new topology is also called the Vietoris topology. Sets of the form Oc(U1, ... ,Uk ) = O(U1, ... ,Uk ) nexpcX, where U1"",Uk are open subsets of X (see the end of §1.3), constitute a base of the new topology. The set expc X with the mentioned topology is called the space of compact subsets of X. If X is a compact urn then by virtue of Theorems 1.6.1 and 1.6.2 expc X = exp X.
Figure 3.3
Figure 3.4
Let now (X, p) be a metric space. In this case consider one more possibility to introduce a topology on the set expc X. For F 1 , F2 E expc X put a(F1 ,F2) = inf{c: c > 0,F2 S;;; O[Fd, see Figures 3.3 and 3.4. Lemma 2.3.6 implies the nonemptiness of the set {c: c > 0, F2 S;;; O[Fd. The formula written defines a number a( Fl , F2)' The number is called the deflection of the set F2 from the set Fl' Notice properties of the deflection
a. For every (nonempty) compact subsets F 1 , F2 and F3 of the space X: 1) a(Fl,F2)~0; 2) a(F1 , F2) = if and only if F2 S;;; F 1 ; 3) a(Fl,F3):::; a(Fl,g) + a(F2,F3). Proof of these properties is simple. Notice only that in the case of the property 3) we need use Lemma 2.3.5. We have not mentioned the symmetry
°
80
CHAPTER 3
between properties of the deflection. The deflection does not possess this property and therefore it is not a metric. The Hausdorff distance between two compact subsets FI and F2 of the space X is defined as h(FI ,F2) = max{a(FI,F2),a(F2,Fd}. In contrast to the deflection the function h is symmetric (i.e.,
By virtue of the above properties of the deflection the function h is a metric on the set expc X (the Hausdorff metric). Theorem 4.1. The Vietoris topology on the space of compact subsets of a metric space is generated by the (corresponding) Hausdorff metric. Proof. Before this theorem is proved we must distinguish the space expc X of compact subsets of the metric space (X, p) with the Vietoris topology and the metric space H(X) with the same carrier set expc X and the Hausdorff metric h. Our aim is to prove that the identity mapping i : H(X) - t expc X is an homeomorphism (see §2.2). Since the bijectivity of this mapping is obvious we have to check the continuity of the mapping i and of its inverse mapping i-I. I. Let U be an open subset of the space X and Fo E Oc(U, X). Fix an arbitrary point x of the intersection Fo n U. Next, fix a number E > 0 such that O~x S;;; U (O~x denotes the E-neighborhood of the x point with respect to the metric p). If F E O~Fo (O~Fo denotes the E-neighborhood of the point Fo with respect to the metric h), then Fo S;;; O~F (the neighborhood with respect to the metric p). By Lemma 2.3.4 there exists a point t of the set F such that p(x, t) < E. SO t E O~x S;;; U. The intersection F n U contains the point t. Hence it is nonempty. Thus i(O~Fo) S;;; Oc(U,X). II. Let U be an open subset of the space X and Fo E 0 c(U). The condition Fo E Oc(U) may be rewritten as Fo S;;; U. By Corollary of Lemma 2.3.8 O,Fo S;;; U for some E > O. By the definition of the Hausdorff metric O~ Fo S;;; O(U), i.e., i(O~ Fo) S;;; O(U). III. The continuity of the mapping i follows from I, II and Theorem 1.7.3. IV. Let Fo E H(X) and E > o. The open cover {0,/4X: x E Fo} of the compactum Fo contains a finite sub cover , = {0,/4XI, ... , Oo/4xd. The following properties of the family, are essential for us:
and (4.2) if points s, t belong to an element of the family" then p( s, t) <
~.
Spaces of mappings and spaces of compact subsets.
81
(the last property follows immediately from the triangle inequality). Lemma 2.3.4 and (4.2) imply immediately that if F I , F2 E O(Or/4XI, ... , Or/4Xk), then h(FI' F2) ::( c/2 < c. By virtue of (4.1) O(Or/4XI, ... ,Or/4Xk) ~ O~ Fo. This means the continuity of the mapping i-I (see Theorem 1.7.3). The theorem is proved. • The same topology on the set X may be generated by different metrics PI and P2· The metrics PI and P2 define (may be, different) Hausdorff metrics hI and h 2 • By virtue of Theorem 4.1 both Hausdorff metrics hI and h2 generate the same topology on the set expc X, namely just the Vietoris topology. The Vietoris topology may be defined directly by the topology of the space X without reference to the metric. The topology of a metric space may be characterized completely by the reserve of convergent sequences, see Theorem 1.5.3. Therefore the just worded independence of the topology of the space expc X on a concrete definition of a metric on the space X may be also deduced from the following assertion. Theorem 4.2. Let F, F 1 , F 2 , • .• be compact subsets of a metric space. The sequence {Fi: i = 1,2, ... } converges with respect to the Hausdorff metric to the compactum F if and only if simultaneously: 1) for every point x E F we can choose points Xi E Fi in a manner, that Xi
~
X,
2) every sequence Xk E Fik (il < i2 < ... ) has a subsequence {Xk : k E A} converging to a point of the set F. Proof. Necessity of condition 1. Denote by Xi a point of the set Fi nearest to the point X (see Lemma 2.3.8. and Figure 3.5). We have
F a· a .>
Figure 3.5
Figure 3.6
Necessity of condition 2. Denote byak a point of the compactum F nearest to the point Xk (Figure 3.6). By virtue of the compactness of the
82
CHAPTER 3
set F the sequence {ak k = 1,2, ... } has a subsequence {ak converging to a point a E F. Then for k E A
P(Xb a) ~ p(Xk, ad
+ p(ak' a)
~ a(F, FiJ
+ p(ak' a)
----t
k E A}
o.
Sufficiency. I. Show that a(F, Fd ----t O. Assume the opposite. There exist a number E > 0 and an infinite set A <:;;; N such that a(F, F k ) ~ E for k E A (Figure 3.7). By virtue of the last inequality there exists a point
Figure 3.7
Figure 3.8
Xk E Fk such that p( x k, F) ~ E. By 2) there exists a subsequence {x k : Ad of the sequence {Xk : k E A} converging to a point x of the set F. Then P(Xk' F) ~ p(Xk, x) ----t 0 as k E Al and k ----t 00. This contradicts the choice of the point Xk: P(Xk' F) ~ E. II. Show that a(Fk, F) ----t o. Assume the opposite. That is, there exist a number E > 0 and an infinite set A <:;;; N such that a(Fk' F) ~ E for k E A (Figure 3.8). By virtue of last inequality there exists a point Xk E F such that P(Xk' Fd ~ E. By virtue of the compactness of the set F the sequence {Xk: k E A} has a subsequence {Xk: k E Ad converging to a point x E F. By 1) we can choose points ak E Fk in a manner that ak ----t x. Then p(ak, xd ~ p(ak' x) + p(x, xd ----t 0, P(Xk' Fd ~ P(Xk' ad ----t 0 as k E Al and k ----t 00. This contradicts the choice of the points Xk: P(Xk' Fd ~ E. III. The sufficiency follows from I and II. The lemma is proved. • Corollary. Let a sequence {FI' F2 , . . . } of compact subsets of a metric space converge with respect to the Hausdorff metric to a compactum F. Let OF be an arbitrary neighborhood of the compactum F. Then there exists an index io = 1,2, ... such that Fi <:;;; OF for i = io, io + 1, .... In the opposite case the set A = {i: i = 1,2, ... , Fqi \ OF -=1= 0} is infinite. For i E A fix a point Xi E Fi \ OF. The sequence {Xi: i = 1,2 ... } has no subsequence converging to a point of set A. This contradicts 2). (However, this corollary may be also deduced from the definition of the Vietoris topology and Theorem 4.1.) •
kE
Spaces of mappings and spaces of compact subsets.
83
Theorem 4.3. The space of nonempty closed subsets of a metric compactum is compact. Proof. I. Let X denote the compact. Let {Fi: i = 1,2, ... } be a sequence of nonempty closed subsets of X. By Theorem 1.5.4 the sequence {Fi: i = 1,2, ... } has a subsequence {Fi: i E A} converging to a set F E expc X in the sense of the fulfilment of the equality
F = lim top inf Fi = lim top sup Fi . i-+oo
i-+oo
The compactness of X implies the nonemptiness of F = lim top sup F i . Let U be nonempty open subset of the compactum X. i--->oo II. Let F E 0 (U). By virtue of Theorem 1.7.2 Fi E 0 (U), beginning with some i = i 1 E A. III. Let F E 0 (U, X). Then Fi E 0 (U, X), beginning with some i = i 1 E A. This is a direct consequence of the definition of the lower limit. IV. Now II and III and Lemma 1.7.3 implies the convergence Fi - 7 F, i E A, with respect to the Vietoris topology. The theorem is proved. •
5. Space of partial mappings In this section we assume that (X, pd and (Y, P2) are metric spaces. Their product X x Y is also a metric space (see §2.1). Let P be the corresponding metric. Every continuous mapping of an arbitrary nonempty closed subspace of the space X into the space Y is called a partial mapping of the space X into the space Y, see also [Za], [Ku2] and [Ku3]. The set of all partial mappings of the space X into the space Y is denoted by Cv(X, Y). Thus Cv(X, Y) = U{C(F, Y): FE expX}. We will also consider partial mappings defined on compact subsets of the space X: Cvc(X, Y) = U{ C(F, Y): F E expc X}. Our next goal is the introduction of a topology and of a metric on the set Cvc(X, Y) which are suitable for the description of properties of solution sets of ordinary differential equations. Introduce a topology on the set Cvc(X, Y) in the following way. Consider the mapping Gr : Cvc(X, Y) - 7 expc(X x Y) of the set Cvc(X, Y) into the space expc(X x Y), associating its graph to a function. The mapping Gr is injective or in other words the mapping Gr : Cvc(X, Y) - 7 Gr (Cvc(X, Y)) is one to one. The set expc(X x Y) carries the Vietoris topology (generated by the corresponding Hausdorff distance). Recall that the topology does not depend on a concrete choice of the metrization of the product (see remarks of the previous section). The set Gr (Cvc(X, Y)) carries a topology and a metric induced from the space expc(X x Y). We define uniquely a
84
CHAPTER 3
topology and a metric on the set Cve(X, Y) when we declare the mapping Gr : Cve(X, Y) --. Gr (Cve(X, Y)) to be an isometry, i.e., we declare the distance between two elements of Cve(X, Y) to be equal to the Hausdorff distance between their graphs. The set Cve(X, Y) with the distance and topology so defined is called the space of partial mappings with compact domains of definition (of the metric space X into the metric space Y). Theorem 5.1. The mapping 1f : Cve(X, Y) --. eXPe X is continuous. Proof. The proof consists in referring to Theorem 3.1 and the continuity of the projection g : X x Y --. X. • For X = JR we have: Corollary. The mappings:
1fb : Cve(JR, Y) --. JR, 1fe : Cve(JR, Y) --. JR, 17r : Cve (JR, Y) --. JR,
1fb(Z) = inf1f(z), 1fe(z) = SUP1f(z), l7r(z) = 1fe(z) - 1fb(Z),
are continuous. • Let H E eXPe X. Denote C:C(X, Y) = {z: Z E Cve(X, Y), H ~ 1f(z)}. For Z E C:C(X, Y) put 0 the set A = {i : i E N, sup{P2(fi(t),f(t)): t E H} ~ c} is infinite. For every i E A fix a point ti E H, such that P2(fi(t i ), f(t i )) ~ E. The mapping cp : H x Y --. Y x Y, cp(t, y) = (f(t), y), is continuous. This follows from Theorem 2.1.1. The mapping 'ljJ : Y X Y --. JR, 'ljJ(x, y) = P2(X, y), is continuous. This follows from Corollary 1 of Lemma 2.3.1. Thus the function a : H x Y --. JR, a(t, y) = P2(f(t), y), is continuous. This implies the openness of the set
U = a- 1 ((-oo,c)) = {(t,y): t E H, y E Y, P2(f(t),y) < c}. in the product H x Y. By the Corollary of Lemma 2.3.8 there exists a number 8 > 0 such that
Spaces of mappings and spaces of compact subsets.
85
0 6 Gr(f) <:;; U u ((X \ H) x Y) = (X
x Y) \ ((H x Y) \ U) (Figure 3.9).
y
Then (ti' f(t i )) ~ 0 6 Gr(f) for every i E A. By virtue of condition 2) of .···~(H Theorem 4.2 and the closedness of the uset (XXY)\06 Gr(f) this contradicts the condition Gr(fi) --t Gr(f). Thus our assumption is false and the theo.:"."'" " : rem is proved. • Theorem 5.3. Under the hypo.X theses of Theorem 5.2 the identity mapping C H (X, Y) --t C(H, Y) is an Figure 3.9 homeomorphism. Proof. The continuity of the mapping follows from Theorem 5.2.
.J
liiii!ol,"~4aL""""";"<"4"'_'
Prove the continuity of the inverse mapping. Let h be the Hausdorff metric of the space Gr(CH(X, Y)). Let hI(f,g) = h(Gr(f),Gr(g)) be the corresponding metric on the space C H (X, Y). Let d be the metric of the uniform convergence in the space C(X, Y), see §1. With this notation we obtain what was required from the obvious estimate hI (f, g) ~ d(f, g) and remarks of §1. 7 about the Lipschitz condition. The theorem is proved. • Theorem 5.4. Let sequences
{Ji: i = 1,2, ... } <:;; Cvc(X, Y) and
{Di: i = 1,2, ... } <:;; expc X converge to f E CvJX, Y) and D E expc X, respectively. Let Di <:;; 7r(fi) (Figure 3.10) for every i = 1,2, .... Then D <:;; 7r(f) and filDi
--t
flD in the space Cvc(X, Y).
Figure 3.10
Proof. 1. The inclusion D <:;; 7r(f) follows from Theorem 5.1, Lemma 1.7.2 and Theorem 3.2.
86
CHAPTER 3
II. Show that the sequence Fi = Gr (JiI D) and the compactum F = Gr (JID) satisfy condition 1) of Theorem 4.2. Let x = (t, f(t)) E F. Since tED and Di ---+ D, by Theorem 4.2 we can fix points ti E Di in a manner that ti ---+ t. Our goal will be achieved w hen we prove the convergence (ti' fi(t i )) ---+ (t, f(t)). If we assume the opposite, we obtain that for some c > 0 the set A = {i: i E N, P((ti' fi(t i )), (t, f(t)))
Figure 3.11
~
c} is infinite. But
(Figure 3.11). Therefore the sequence {(ti, fi(t i )): i E A} contains a subsequence {(ti,fi(ti)): i E Ad converging to a point y = (8,1(8)) of the set Gr(f). By Lemmas 2.1.1 and 2.1.2 ti ---+ 8 for i E Al and i ---+ 00. Since ti ---+ t, t = 8. Therefore y = (t, f(t)). This contradicts the definition of the set A. III. Show now that the sequence Fi = Gr (JdD) and the compactum F = Gr (JI D) satisfy condition 2) of Theorem 4.2. Let a set A S;;; N be infinite. Let ti E Di and Xi = (ti' fi(t i )) for i E A. By virtue of the convergence Di ---+ D and Theorem 4.2 the sequence {t i : i E A} contains a subsequence {t i : i E Ad converging to a point t of the set D. Since (ti,fi(t i )) E Gr(fi) ---+ Gr(f), the sequence {(ti,fi(t;)): i E Ad contains a subsequence {( t i , fi (t i )): i E A 2 } converging to a point y = (8, f(8)) of the set Gr(f). By Lemmas 2.1.1 and 2.1.2 ti ---+ 8 as i E A 2 , i ---+ 00. Since ti ---+ t, 8 = tED. Therefore the subsequence {( t i , fi (t i )): i E A 2 } converges to the point y = (t, f(t)) of the set F = G(fID)' This is just what was • required. The theorem is proved. The position this theorem is well clarified by: Corollary. Let X be a compactum and Y be a Hausdorff space. Then the mapping : C(X, Y) x X ---+ Y, (z, x) = z(x), is continuous. To obtain the corollary let us notice that by virtue of Theorem 5.3, Theorem 5.4 implies the equality lim(s,t)-+(z,x) (8, t) = (z, x) for every point (z, x) of the product C(X, Y) x X (see also Lemma 1.7.2). It means just the continuity of the mapping at the point (z,x). •
Spaces of mappings and spaces of compact subsets.
Theorem 5.5. Lei {Ii: i = 1,2, ... } ~ CVC(X, Y), f E Cvc(X, Y), {Di i 1,2, ... } U {Ei i = 1,2, ... } ~ expc X. Lei D, E E expc X, Di U Ei ~ 7r(fi) (Figure 3.12) for every i = 1,2, ... , DUE ~ 7r(f), filDi - t fiD' hlEi - t fiE· Then
filDiUEi
-t
87
-E
fI DUE .
x
Proof. The theorem will Figure 3.12 be proved, when we check the fulfilment of conditions 1) and 2) of Theorem 4.2 for the sequence of the compacta Fi = Gr (JiIDiUE) and for the compactum F = Gr (JI DUE ). I. Take arbitrarily x = (t,f(t)) E F = Gr (JI DUE ). Let, tED for definiteness' sake. By virtue of the convergence filDi - t flD and by Theorem 4.2 we can fix points Xi E Gr (JiI D ) in a manner that Xi - t x. Thus Xi E Gr (JiI DiUE ) and Xi - t x. This means the fulfilment of condition 1). II. Let a set A ~ N be infinite, ti E Di U Ei and Xi = (t i , fi(t i )) for i E A.
Since A = Al U A 2, where
Al={i: iEA, tiEDi}, and A 2 ={i: iEA, tiEE;}, at least one of these two sets Al or A2 is infinite. Let the set Al be infinite for definiteness. By virtue of the convergence Gr (JiI D ) - t Gr (JI D ) and by Theorem 4.2 we can pass to a subsequence {(ti,fi(t i ))· i E B} of the sequence {(ti,fi(t;)): i E Ad converging to a point X of the set Gr (JI D ) ~ Gr (JIDUE). This means the fulfilment condition 2). The theorem is proved. •
6. Compactness in the space of partial mappings A subset Z of the set Cv(X, Y) of partial mappings of a metric space (X, pd into a metric space (Y, P2) is called equicontinuous, iffor every number c > 0 there exists a number 0 > 0 such that P2(Z(S), z(t)) < c for every mapping z E Z and for every two points s, t E 7r(z), which satisfy the condition Pl(S, t) < o. ~
88
CHAPTER 3
Theorem 6.1. Let (X, pd and (Y, P2) be metric spaces. Then every compact subset of the space Cvc(X, Y) is equicontinuous. Proof. Under these hypotheses the space Cvc(X, Y) is metrizable. Let Z be a compact subset of the space Cvc(X, Y). The definition of the Vietoris topology implies immediately that the mapping Gr (as a multi-valued mapping into the product X x Y) is upper semicontinuous. Theorem 1. 7.8 implies the compactness of the set E = U{Gr(z): z E Z}. Assume that the set Z is not equicontinuous. This means, in particular, that for some E > 0 for every k = 1,2, ... there exist a function Zk E Z and points Sk, tk E 7r(Zk) such that Pl(Skl td < 2- k and p2(zdsd, zdt k )) ~ E. By virtue of the compactness (and the metrizability) of the set Z we can assume in addition that the sequence {Zk: k = 1,2, ... } converges to a function Z E Z. By virtue of the compactness of the set E we can assume in addition that the sequences {( s k, Zk (s k)): k = 1, 2, ... } and {(t kl Zdtk)): k = 1,2, ... } converge, respectively, to points (s, yd and (t,Y2). Theorem 5.4 implies the equalities Yl = z(s) and Y2 = z(t). The estimate PI (s, t) ~ PI (s, sd + PI (Skl t k ) + PI (tk' t) follows from the triangle inequality. All terms in its right hand side tend to zero as k - t 00. Therefore PI(S,t) = O,s = t, z(s) = z(t). On the other hand,
When we pass to the limit as k - t 0 in the right hand side of the inequality we obtain the estimate P2(Z(S), z(t)) ~ E. Thus our assumption (that the set Z is not equicontinuous) has led to a contradiction. The theorem is proved. • The theorem proved remains valuable in the case when the set consists of one function only. Usually we use here another terminology and we call a function which constitutes an equicontinuous family uniformly continuous. Thus we obtain: Corollary. A continuous function defined on a metric compactum is uniformly continuous. • Theorem 6.2 (Arzela). Let K be a compact subset of the product X x Y of metric spaces (X,pd and (Y,P2). Let the sequence 0: = {Zk : k = 1,2, ... } ~ Cvc(X, Y) be equicontinuous. Let Gr(zk) ~ K for every k = 1,2, .... Then the sequence 0: has a convergent (in the space Cvc(X, Y)) subsequence. Proof. By virtue of Theorem 4.3 we can assume in addition the convergence of the sequence {Gr(zk): k = 1,2, ... } in the Vietoris topology to a point F of the space expc K. Denote by g the restriction to the set F of the projection X x Y - t X. Show that the mapping g : F - t X is injective. Assume the opposite, i.e.,
Spaces of mappings and spaces of compact subsets.
89
that there are two different points (X,YI) and (X,Y2) of the sets F with coinciding first coordinates. Put c = ~P2(YI' Y2) (> 0). By Theorem 4.2 we can choose points Sk, tk E 7r(zd for k = 1,2, ... in a manner that (Sk' ZdSk)) -- (x, yd and (tk' zk(td) -- (x, Y2)' By Lemma 2.1.2 Sk -- x and tk -- X. By virtue of equicontinuity of the sequence 0: there exists a number 8 > 0 such that p2(zds), zdt)) < c for every index k = 1,2, ... and every points s, t E 7r(zd satisfying the condition PI(S, t) < 8. Since Sk -- x and PI (s kl
x) <
8
2"'
beginning with some k. Hence:
(6.1) By the triangle inequality:
By Lemma 2.1.2 the first and third terms in the right hand side of (6.2) tend to zero as k -- 00. Because of (6.1)
Our assumption has led to a contradiction. So the mapping 9 is injective. This means that the set F is the graph of a mapping z : g(F) -- Y. By Theorem 2.2.1 the set g(F) is compact and the mapping h = g-l (inverse to 9 : F -- g(F)) is continuous. The mapping z may be represented in the form of the composition of the mapping h and the projection X x Y __ Y. This implies the continuity of z. Thus z E Cvc(X, Y). The convergence Gr( Zk) -- Gr( z) and the definition of the topology of the space of partial mappings imply the convergence Zk -- z. The theorem is proved. • 7. Space C.(M)
Let M be a subset of the product lR x X of the real line lR and of a metric space X. Denote by C.(M) the set {z: z E Cvc(lRxX), 7r(z) is connected, Gr(z) ~ M}. The topology induced on the set C.(M) by the space of partial mappings Cvc(lR, X) makes the set C.(M) a topological space. By the remarks of §5 this space is metrizable. L
90
CHAPTER 3
By virtue of Theorems 2.6.3 and 3.3 the set of all segments is closed in the space expc JR. Therefore the continuity of the mapping 7f implies the closed ness of the subspace Cs(JR x X) in the space Cvc(JR, JRn). So Theorem 6.2 implies: Theorem 7.1. Let K be a compact subset of the product JR x X, a be an
equicontinuous sequence of elements of the set Cs(K). Then the sequence a has a convergent (in the space Cs(K)) subsequence. • In the rest part of this section we assume that X = JRn. Theorem 7.2. The mapping I: Cs(JR x JRn) --+ JRn,
J b
I(z) =
z(t)dt,
a
where [a, b] = 7f(z), is continuous. Proof. Take an arbitrary function Zo E Cs(JR x JRn) and an arbitrary number E > o. Let 7f(zo) = lao, bolo Let a number M > 0 be such that Im(zo) ~ 0(0, M) (see Theorem 1.7.8 and Lemma 2.3.6). By virtue of the continuity of the mappings 7f and 1m (see Theorem 3.1) there exists a neighborhood VI of the point Zo in the space C s (JR x JRn), for every element z of which -
Im(z) ~ 0(0, M), lao - al
E
E
< 8M and Ibo - bl < 8M'
where [a, b] = 7f(z). If ao = bo, then for z E VI
J
J
J
ao
a
a
bo
zo(t)dt -
Let ao
b
b
2E
z(t)dt :::;; 8M . M <
z(t)dt
E.
<
bo. Select points al and bl in order to satisfy the inequalities ao < al < bl < bo, lao - all < 8~' Ibo - bll < 8~· By virtue of the continuity of the mapping 7f there exists a neighborhood V2 of the point Zo in the space C s (JR x JRn) such that a < al and bl < b for every function z E V2 , where as above [a, b] = 7f(z). For z E VI n V2 we have a, b, b bo
j zo(t)dt - j z(t)dt :::;; j(zo(t) - z(t))dt a, a ao
j zo(t)dt b,
j zo(t)dt ao
a,
bo
+
+
+
j z(t)dt a
b
+
j z(t)dt b,
Spaces of mappings and spaces of compact subsets.
91
bt
~
j(zo(t) - z(t))dt
<
j (zo(t) - z(t))dt
bt
3c
+ 4'
Theorems 5.3 and 2.1 imply the existence of a neighborhood V3 of the point Zo in the space Cs(lR, IRn) such that Ilzo(t) - z(t)11 < 4(bl~a,) for every function z E V3 satisfying the condition 7r(z) 2 [aI, bd and for every point t E [aI, bd· Hence for z E VI n V2 n V3 bt
j(zo(t) - z(t))dt
~
al
bt
j Ilzo(t) - z(t)lldt al
III(zo) - I(Z)II <
c
3c
4+ 4
= c.
This means the continuity of the mapping I (at an arbitrary point Zo E Cs (IR x IRn) and hence on the entire space Cs (IR x IRn)). The theorem is proved. • Theorem 7.3. With the notation of Theorem 7.2 let 1
J(z) = b _ a I(z), if b =1= a and
J(z) = z(a),
= b. Then the mapping J : C s (IR x IRn) - t IRn is continuous. Proof. The continuity of the mapping J at a point Zo E Cs (IR x IRn), l,.(zo) > 0, follows immediately from Theorem 7.2, from the continuity of the mapping 1,. (see Corollary of Theorem 5.1) and of the division of vectors by nonzero numbers. Let 7r(zo) = {ao}. Let z E Cs(JR x JRn) and 7r(z) = la, b]. For a = b: if a
IIJ(zo) - J(z)11 = IIzo(ao) - z(a) II
~
h(Gr(zo), Gr(z)),
CHAPTER 3
92
where h denotes the Hausdorff metric generated by the Euclidean metric of the space IR x IRn = IRn+l. For a f- b:
IIJ(zo) - J(z)11
~ b~ a
b
J J
Ilzo(ao) - z(t)lldt
a
~ b~ a
b
h(Gr(zo), Gr(z))dt
a
= h(Gr(zo), Gr(z)).
These estimates imply the continuity of the mapping J at the point Zo. The theorem is proved. •
8. Alexander's lemma and its consequences Zorn's lemma 1.8.1 allows us to obtain results of wide generality. With the same success we may use assertions related to the axiom of choice and being equipotent to Zorn's lemma. However, t.he discussion of this question lies a long way from the framework of this book. In this section we will notice only possibilities for clarifying the meaning of methods used and some relations. We will generalize Theorems 2.2.2 and 4.3 with the help of Zorn's lemma. (In the main part of our account the theorems was not proved in their maximal generality because the corresponding methods are not usual in the circle of questions under consideration. It is sufficient to have versions of them obtained in the development of the technique of reasoning which will be necessary anywhere in the further account.) Lemma 8.1 (Alexander). Assume that a topological space X possesses a sub-base (3 such that: every cover of the space X by elements of (3 has a finite subcover. Then the space X is compact. Proof. Our aim is to show that
every open cover, of the space X has a finite sub cover. Assume the opposite. Let r denote the set of open covers space X without finite subcovers. Our assumption means that r is nonempty. Let
,0
E
r.
1. The set r is partially ordered by the relation ~. Lemma 1.8.2 implies the existence of its maximal linearly ordered subset N 3 {'o}.
Spaces of mappings and spaces of compact subsets.
93
II. If {'Yl, ... , Ik} ~ N then we can renumber the set 11, ... "k according to its order. That is, we can assume in addition the fulfilment of the condition 11 ~ ... ~ Ik' Then ubi,"" Id = Ik EN. III. Show that the cover I' = uN belongs to N. Assume the opposite, that is, that the cover I' has a finite subcover {U1 , ... , Ud. For every i = 1, ... ,k fix a cover Ii E N containing Ui . Since bl,' .. "d ~ N, by II u{ 11, ... "k} = Ik E N. By our choice the cover Ik has a finite sub cover {U1 , • •• ,Ud. This contradicts the condition Ik E N. IV. Show that the family f3' = I' n f3 covers the space X. If it is not so, there exists a point x E X \ (Uf3·). An element U of the cover I' contains the point x. Fix elements Vi,' .. , Vm of the sub-base f3 such that x E n {Vi, ... , Vm} ~ U. By virtue of the choice of the point x none of the sets Vi,' .. , Vm belongs to I'. By virtue of the maximality of the cover I' the cover {Vi} U I" i = 1, ... , m, has a finite sub cover
{Vi, W{, ... , W~(i)}' We have:
.......
and n{V1 , ... , Vm } ~ U. The finite subfamily {U} U {W/: j = 1, ... ,m, i = 1, ... ,n(j)} of I' covers the space X. This contradicts the condition 1* E f. Thus the additional assumption is false and the lemma is proved .• Theorem 8.1 (Tychonov). The product X = IT {Xa: a E A} of compact spaces X a , a E A, is compact. Proof. Sets of the form p;;l(U) (the notation of §2.1), where U runs over the topology of the space Xa and a runs over the set A, constitute a sub-base f3 of the Tychonoff topology of the product X = IT {Xa: a E A}. In order to refer to Alexander's lemma it is sufficient to check that an arbitrary cover of the space X by sets from f3 has a finite subcover. Assume the opposite. Let a cover I ~ f3 of the space X has no a finite subcover. For a E A denote by la the set of all elements of I which can be represented as p;;I (U) (just with this a). If for some a E A the family I~ = {U : U ~ X a , p;:;I(U) E la} covers the space X a , then by virtue of the compactness of the space Xa the cover I~ has a finite sub cover I~' This means that the cover I has a finite sub cover {p;;l(U): U E I~}' which contradicts the initial assumptions about the cover I' It remains to consider the case when for every a E A the family I~ does not cover Xa. In this case we have a possibility associate to every a E A a point Xa E Xa \ U/~, The point {xa: a E A} E X is not covered by,. By the initial assumption I covers X. Thus our additional assumption is false and the theorem is proved. •
94
CHAPTER 3
Theorem 8.2. If a space X is compact, then the space exp X is compact too.
Proof. Sets of the forms 0 (U) and 0 (U, X), where U runs over the topology of the space X, constitute a sub-base of the Viet oris topology of the space exp X. Let I be an arbitrary cover of the space X by elements of the mentioned subbase. Show that the cover I contains a finite subcover. Let Y = X\U where II is the set of all open subsets U of the space X such that 0 (U, X) E I. If Y = 0 then the family II covers the space X and by virtue of the compactness of X II has a finite subcover 12. Then exp X = U{ 0 (U, X) : U E 12}. So the cover I has a finite subcover. If Y =1= 0 then the point Y of the space exp X does not belongs to any set 0 (U, X) E I. Since I is a cover of the space exp X the point Y belongs to one of sets 0 (U) E I. Then Y ~ U. The cover II of the set X \ U contains a finite sub cover {U1 , ... , Urn}. It remains to notice that the family {O (U) ,0 (U1 , X) , ... ,0 (Urn' X)} covers the space exp X: if a point F E exp X does not belong to 0 (U1 , X) U ... U 0 (Urn' X), then F n (U: 1 Ui ) = 0 and therefore F ~ U, FE 0 (U). Referring to Alexander's lemma completes the proof. •
,1,
CHAPTER 4
DERIVATION AND INTEGRATION
Below we develop a theory directed to the study of properties of solutions of ordinary differential equations y' = f (t, y). It will be helpful in the investigation of equations with complicated discontinuities in their right hand sides. A natural approach to the definition of a solution of an equation is to consider a function as a solution, if at every point of its domain the function has a derivative satisfying the equation. This plan may be realized, for instance, when we study equations with continuous right hand sides. But there are many very simple cases when solutions in the sense stated are absent; however, there exist solutions in some near but different meaning. Moreover, these solutions correspond to the contents of the (physical, geometric, etc.) problem solved with the help of the equation under consideration. Our aim in this chapter is to introduce and to discuss a notion of the derivative which suits for large spectrum of such situations. A natural restriction is that if a problem may be solved with the help of classical methods then no new 'solution' appears on account of the play on definitions. As Integration theory in general and the theory of Measurable Mappings lie outside the framework of the theory of Ordinary Differential Equations, we will not try to give them here a complete account. We will follow the shortest way to a limited list of results and facts important for the understanding of the main contents of the book. It seems this is not a drawback. 1. Measure of a set on the line
Let M be a subset of the real line R When we cover the set M by an at most countable family of intervals 'Y = {(ak' bk ) k = 1,2, ... } then it is natural to say that b2 the 'length' of the set M 1---+-) --1(-+-)-+-)(---+does not exceed the sum d( 'Y) of the lengths (== total length) of intervals of the family 'Y, see FigFigure 4.1 ure 4.1. This argument pushes us to the consideration of the lower bound j.L*(M) of numbers db), t--(
96
CHAPTER 4
where, runs over the set H(M) of all at most countable families of intervals covering the set M. The word 'numbers' is used here conditionally: it may be, for all, E H(M) the sum d(,) is infinite. In addition, we may assume that the family, may contain intervals of infinite length. In these cases we write d(,) = 00. The number J.t*(M) (under the same reservation: it may be that J.t * (M) = (0) is called the outer measure of the set M. Lemma 1.1. If Ml ~ M2 (~~), then J.t*(Md ~ J.t*(M2). Proof. The proof is obvious: we obtain immediately what was required • from the inclusion H(M2 ) ~ H(Ml). Lemma 1.2. Let {Mk : k = 1,2, ... } be a family of subsets of the real line. Then
Proof. When the outer measure of one of the sets M k , k = 1,2, ... , is infinite or when the sum in the right hand side is infinite the proof is obvious. In what follows we assume the opposite. Take an arbitrary E > o. For every k = 1,2, ... fix a family of intervals such that U, "2 Mk and k d('d ~ J.t*(Mk) + 2- E. Such a family exists by the definition of the outer measure. For the family, = U{ ,k: k = 1,2, ... } we have
'k
U {Mk: k = 1, 2, ... } ~ d(,)
= 2:)d('d:
k
u"
= 1,2, ... }
~L{J.t*(Md+2-kc: k=1,2, ... } 00
k=l
=
L{J.t*(Md: k
= 1,2, ... } +E.
Hence J.t*(U{Mk: k = 1,2, ... }) ~ {J.t*(Md: k = 1,2, ... } + c.
Since this is true for every number c > 0 the lemma is proved. • The outer measure of the empty set is equal to zero. In Lemma 1.2 we did not assume the nonemptiness of the sets M k • Therefore the analog of the inequality just proved for a finite family of sets is its particular case. Lemma 1.3. Let a subset of the real line be represented as a union of a finite number of pairwise disjoint segments. Then the outer measure of the subset is equal to the total length of these segments. Proof. Let M = U{[ak, bkJ: k = 1, ... , s} be the set and its representation from the statement of the lemma. Notice that by virtue of
97
Derivation and integration.
Theorem 2.6.7 these segments, and only they, are connected components of the set M and such a representation is unique. Let d = 2:~=1 Ib k - ak I. The inequality J.L*(M) ~ d follows from the consideration of the families = {( ak - E, bk + E): k = 1, ... , s}, E > O. The lemma will be proved when we show that d(,) ~ d for every (at most countable) family, of intervals covering the set M. By virtue of the compactness of the set M it is sufficient to prove this assertion under the assumption of the finiteness of the cover ,. It will be done with the help of an induction on the number of elements of, = {U1 , . . . , Ud. We will do it not for a fixed set M, but for all such sets at once. For t = 1 the assertion is obvious. Let the assertion be proved for t = i - I ~ 1. Prove it for t = i. We do not lose any generality when we assume in addition that the segments [ak, bd are enumerated according to their position in the line: al ~ b1 < a2 ~ b2 < ... < as ~ bs . We can also assume that just the interval Ut = (p, q) contains the point bs (in the opposite case we renumber the family,). If p < al then as in
,t:
V1
Vt
V3 ~
~
~
( ( ) () ) ........ ".... ( -+ -t-+---t----1f-+-t-at
() ( )
}
b1 a2 '-..--'
V2 Figure 4.2
......
the case t = 1, the assertion is obvious. Consider the case al ~ P « bs ) (Figure 4.2). Let Ml = M n [al,pl and M2 = M n [p, bsl. Each of these sets may be also represented as the union of a finite number of pairwise disjoint segments. If d1 and d2 are the total lengths of the components of Ml and M 2, respectively, then d = d 1 + d 2. The family,1 = {U1 , •.• , Ut-d covers the set MI. By the inductive hypothesis d(,I) ~ d 1 • The length of the interval Ut is not less than d 2. So d(,) = d( '1) + d( {Ut }) ~ d 1 + d 2 = d. The lemma is proved. • Lemma 1.4. Let an (at most) countable family, = {Mk : k = 1,2, ... } consist of pairwise disjoint segments, intervals and half intervals of the real line and E > O. Then: a) if d(,) < 00, then there are segments II"'" I j lying in different elements of the family, such that: J.L*(I1 U··· U I j ) ~ d(,) - c and J.L*«U,) \ (U{Il,"" Ij}) < E; b) if d(,) = 00, then there are segments I 1 , ••• ,Ij lying in different elements of the family, such that J.L*(I1 U ... U I j ) ~ c .
98
CHAPTER 4
Proof. Let ak ~ bk be endpoints of a segment, an interval or an half interval M k , k = 1,2, .... If ak = -00 or bk = 00 for some k, then we have b) and the validity of the assertion is obvious: the set Mk contains segments of arbitrarily large length. Below we assume that -00 < ak ~ bk < 00. For k = 1,2, ... denote by h the segment [ak + 2- k - 3 E, bk - 2- k - 3 E], if 2- k- 2E ~ bk - ak' If the last condition is not fulfiled we put h = 0. The length l(Id of the segment (or empty set) h is not less than the number max{bk - ak - 2- k- 2E,0}. For the calculus of d(,) we have the series L~=1 (b k - ad· For the calculus of the total length of the segments h, k = 1,2, ... we have the series L~=1 l(h). Let us estimate the difference of partial sums of these series: i
i
L
i
2)bk - ad l(h) k=1 k=1
~
L
Ibk - ak -l(h)1
k=1
~
i
L 2- k- 2
E
< ~.
k=1
Therefore the sums of these series, i.e., the limits of the sequences of its partial sums, differ at most by ! < ~, if db) < 00, and coincide if db) = 00. There exists a number j such that l(Id+· .. +l(Ij) > db) - ~ (> db) -E), for db) < 00 and l(Id + ... + l(Ij) > 10 for db) = 00. By Lemma 1.3 the last observation completes the proof in the case b). In the case a) the previous estimate proves the first part of the assertion. The second part of a) follows from the fact that difference (U,) \ (U{ II, ... ,Ij}) lies in the union of the sets [aI, al + 2- 4 10], [b 1- 2- 4 10, b1], ... , [aj, aj + 2- j - 3 E], [bj - 2- j - 3 E, bj ], [aj+l' bj+l], [aj+2' bj+2], .... Therefore Lemmas 1.1 and 1.3 and the inequality
imply the estimate
fL*((U,) \ (U{Il,'" ,Ij}) ~ 2- 3 10 + ... + 2- j - 2E + (bj+l - aj+l) + (bj+2 - aj+2)
<
J+
db) - (db) -
+ ...
~)
<
E.
The lemma is proved. • Lemma 1.5. Let G be an open subset of the real line and 10 > 0. Then there exists a closed subset H of the real line such that H ~ G and fL*(G \ H)
< E.
Proof. Theorems 2.8.8 and 2.8.3 guarantees the representation of an open subset of the line as the union of an at most countable family of disjoint intervals. Use Lemma 1.4 to find for every k = 0, ±1, ±2, ... a closed subset Mk of the line such that
99
Derivation and integration.
Mk t;;;; G n (k, k + 1) and J-L*((G Let M =
{a, ±1, ±2, ... } n G.
n (k, k + 1)) \ Md <
2- lkl - 2 E.
The set
H = M U (U{Mk: k = 0, ±1, ±2, ... }) is closed in the line. (Every point x E IR lies in an interval of the form (k, k + 2). If x rf. H then x rf. M U Mk U Mk+l. There exists a neighborhood Ox t;;;; (k, k + 2) of the point x without points of the set M U Mk U Mk+l. Since Ox n H = 0, x rf. [H]. In view of the arbitrariness of the point x E IR \ H this means the closedness of the set H.) By Lemma 1.2 00
00
k=-oo
k=O
The lemma is proved. • Lemma 1.6. Let F be a closed subset of the real line and 10 > 0. Then there exists a neighborhood OF of the set F such that J-L * (0 F \ F) < E. Proof. For G = IR \ F let us find H according to Lemma 1.5. When we take OF = IR \ H we obtain what was required because OF \ F = G \ H. The lemma is proved. • A subset M of the real line is called measurable if for every 10 > there are a closed subset F and an open subset G (Figure 4.3) of the line such that F t;;;; M t;;;; G and J-L * (G \ F) < E.
°
--+----+ G: f-
-)
+------ + F:
)-
.
H: Figure 4.3
Lemmas 1.5 and 1.6 imply that open and closed sets are measurable. Evidently the complement to a measurable set is measurable (see the proof of Lemma 1.6). If a set M t;;;; IR is measurable its outer measure J-L*(M) is called the measure of the set M and is denoted by J-L(M). Lemma 1. 7. The intersection of two measurable subsets of the real line is measurable. Proof. Let M i , i = 1,2, be measurable subsets of the line and c > 0. Take a closed Fi and an open G i sets such that Fi t;;;; Mi t;;;; G i and J-L(G i \ F i ) < ~. Evidently Fl nF2
C;;;;
Ml nM2 t;;;; G l nG 2
100
CHAPTER 4
and By Lemma 1.2
/L*((GI n G 2 )
\
(FI n F2 ))
~ /L*(GI \ FI) + /L*(G 2 \ F2 ) < ~ + ~
=
E.
The lemma is proved. • Lemma 1.7 and the above remark about the measurability of a complement imply: Lemma 1.8. The difference of two measurable subsets of the real line is measurable. • In addition, MI U M2 = IR \ ((1R \ M I ) n (IR \ M 2)) and Lemma 1.7 and the same remark (or Lemma 1.8) imply the measurability of the union of two measurable set. An obvious induction on the number of elements gives: Lemma 1.9. The union of a finite family of measurable subsets of the real line is measurable. • Lemma 1.10. Let MI and M2 be disjoint closed subsets of the real line. Then ./L(MI U M 2 ) = /L(MI) + /L(M2 ).
Proof. The inequality /L(MI U M 2 ) ~ /L(MI ) + /L(M2 ) follows from Lemma 1.2. Our lemma will be proved when we establish the validity of the inverse inequality. Let, be an arbitrary cover of the set MI U M2 by intervals. By virtue of the normalness of the line there exist disjoint neighborhoods OMI and OM2 of the sets MI and M 2 , respectively. The sets Vi = (U,) n OMI and V2 = (U,) n OM2 are also disjoint neighborhoods of the sets MI and M 2 . The set of connected components of the set V = VI UV2 is split in two subset: components of the set Vi and components of the sets V2. By Lemmas 1.2-4 the measure of an open set is equal to the total length of its connected components. Therefore /L(V) = /L(Vd + /L(V2 ). By Lemma 1.2 /L(U,) ~ d(r). By Lemma 1.1 /L(V) ~ /L(U,) and
/L(Md
+ /-l(M2 )
~
/L(Vi)
+ /L(V2).
Thus
In view of the arbitrariness of the cover, we have
The lemma is proved.
•
101
Derivation and integration.
Theorem 1.1. Let measurable subsets M k , k = 1,2, ... , of the line be pairwise disjoint and L~l J.t(Md < 00. Then the set U{Mk : k = 1,2, ... } is measurable and 00
J.t(U{ M k : k
= 1,2, ... }) = L
J.t(Mk )·
k=l
Proof. Take an arbitrary c > O. According to the convergence of the series L~l J.t(Md fix an index j such that
Now Lemma 1.2 implies the existence of a family of intervals, such that
U{Mk: k = j
+ l,j + 2, ... } ~ U, and db) < ~.
For k = 1, ... , j fix a closed Fk and an open G k sets such that ~ Mk ~ G k and J.t(G k \ Fk ) < 2- k - 1 c. The sets F = FI U ... U Fj and G = G 1 U ... U Gj U (U,) satisfy the conditions
Fk
F ~ U{Mk: k = 1,2, ... } J.t(G \ F) ::::; J.t(G I \ Fd + ... + J.t(G j ::::;
c + ...
2- 2
+ 2-
In view of the arbitrariness of c
j - 1
c+
~ \
2- l
G, Fj ) c
+ J.t(U,)
< c.
> 0 this means the measurability of the
set
M = U{Mk: k = 1,2, ... }. By Lemmas 1.1 and 1.10 j
00
L J.t(Md k=l
c ::::;
L J.t(Mk) k=l
c
j
2 ::; L
J.t(Fk) = J.t(F)
k=l
::::; J.t(M) ::::; J.t(G) ::::; J.t(F) + J.t(G \ F) ::::; J.t(F)
k=l
k=l
In view of the arbitrariness c > 0 00
J.t(M) =
L
k=l
J.L(Mk).
+c
102
CHAPTER 4
The theorem is proved. • Lemma 1.11. A set M C ~ is measurable if and only if for every segment I ~ ~ the set M n I is measurable. Proof. Necessity follows from Lemma 1.7. Sufficiency. Take an arbitrary c > 0. For every k = 0, ±1, ±2, ... fix a closed Fk and an open G k subsets of the line such that
The set F = U{Fk: k = 0, ±1, ±2, ... } is closed. (Every point x E ~ lies in an interval (k, k + 2). If x f/. F, then x f/. Fk U Fk+I' The set Ox = (k, k+2)\(FkUFk+d is a neighborhood of the point x without points of the set F.) The set G = U{G k : k = 0, ±1, ±2, ... } is open and 00
F
~
M
~
J1.(G \ F) ~
G,
L k=-oo
00
2- lkl - 2 c
< L 2- k- I c = c. k=O
The lemma is proved. • Theorem 1.2. Let {Mk: k = 1,2, ... } be a family of measurable subsets of the real line. Then the set M = U{Mk: k = 1,2, ... } is measurable and
Proof. The proof use Lemma 1.11. Let [a, b] be an arbitrary segment of the line. We have the representation M n [a, b] = where
U"
, = {(Mk
n [a, b]) \
(U{MI , ... , Mk-d): k = 1,2, ... }.
The measurability of elements of the family , follows from Lemmas 1.8 and 1.9. By Theorem 1.1 and Lemmas 1.1 and 1.3 partial sums of the series l: {J1.(r) : r E ,} are bounded from above by the number b-a. Therefore the series converge. Theorem 1.1 implies the measurability of the set M n [a, b]. Referring to Lemma 1.11 completes the proof of the measurability of the set M. The validity of the inequality follows from Lemma 1.2. The theorem is proved. • Assertion 1.1. Under the hypotheses of Theorem 1.2 let MI ~ M2 ~ M3 ~ .... Then J1.(M) = .lim J1.(Mi)' '-+00 Proof. I. Lemma 1.1 implies that:
II. If limi-+oo J1.(Mi ) =
00,
then by I the fact in question is obvious.
103
Derivation and integration.
III. Let limi ..... oo J.t(Mi ) <
00.
By Theorem 1.1:
J.t(M) = J.t(M1 ) + J.t(M2 \ Md + J.t(M3 \ M 2 ) + ... = 1-+00 .lim (J.t(Md + J.t(M2 \ M 1 ) + ... + J.t(Mi \ Mi-d) = lim J.t(M;). 'l.~OO
The assertion is proved. • Theorem 1.3. Let {Mk: k = 1,2, ... } be a family of measurable subsets of the real line. Then the set n{Mk: k = 1,2, ... } is measurable. Proof. By virtue of the equality
n{Mk: k=1,2, ... }=~\(U{~\Mk: k=1,2, ... }) our assertion follows from Lemma 1.8 and Theorem 1.2. The theorem is proved. • Assertion 1.2. Let under the hypotheses of Theorem 1.3 J.t(Md < 00 and Ml ;2 M2 ;2 M3 ;2 . . .. Then
Proof.
J.t(M) = J.t(Md - J.t(M1 \ M) = J.t(Md - J.t(U{ Ml \ M i ~ J.t(Md - .lim J.t(M1
\
:
i = 1,2, ... })
Mi )
' ..... 00
= lim (J.t( Md - J.t( Ml \ M i )) ' ..... 00
The equality (*) is true by Assertion 1.1. The assertion is proved. • Example 1.1. A set M satisfying the condition J.t*(M) = 0 is called a set of measure zero. Every such a set is measurable (for the verification of the condition occurring in the definition of a measurable set we can take F = 0). Lemma 1.2 implies that the union of an (at most) countable family of sets of measure zero is a set of measure zero. In particular, the measure of an one point set is equal to zero. So the measure of every countable set is equal to zero (in particular, the set is measurable ). Example 1.2. Calculate the measure of the Cantor perfect set described in Example 1.6.4. Use the notation introduced there. For an arbitrary segment [a, b] we have J.t(S([a, b])) = ~J.t([a, b]). So J.t(Sk+l([O,l])) = ~J.t(Sk([O, 1])) and J.t(Sk([O,l])) = (~)k for every k = 1,2, .... By Lemma 1.1 J.t(K) ~ (~)k for every k = 1,2, .... Hence J.t(K) = O.
104
CHAPTER 4
Let M be a measurable subset of the line. A subset Mo of the set M is called a subset of full measure if the measure its complement in M is equal to zero: /l(M \ Mo) = O. The complement in the segment [0, 1J to the Cantor perfect set gives an example of a set of full measure. The set of rational numbers is countable. Hence its measure is equal to zero. Therefore the set of irrational numbers is a set of full measure on the line. rt is suitable also to use the following terminology. Let P be a property such that points of a measurable set M can possess or not possess it. We will say that the property P is fulfilled almost everywhere on the set M (or at almost all points of the set M) if the set of all points of the set M, in which the property P holds, is a subset (of the set M) of full measure. Non-measurable subsets exist. Here we will abstain from the construction of a corresponding example. 2. Measurable functions A function (a mapping) f defined on a measurable set M ~ ~ and taking values in a topological space Y is called measurable if for every open subset U of the space Y the set f-l(U) is measurable. By Theorem 1.3.3 and Lemma 1.8 a function f measurable if and only if for every closed subset H of the space Y the set f-l(H) is measurable. We say that a function f is measurable on a set M ~ 7r(J), if the function 11M is measurable. Example 2.1. A continuous function is measurable (see §1 and Theorem 1.7.5). Let functions fl : Ml --+ Y and 12 : M2 --+ Y be defined on measurable sets Ml and M2 and take values in a set Y. We say that the functions II and 12 coincide almost everywhere, if
We say that the functions II and f2 coincide almost everywhere on a measurable set M ~ ~ if the functions flIMn7l"(h) and f2IMn7l"(h) coincide almost everywhere. Lemma 2.1. Let functions fl and 12 be defined on measurable sets Ml and M 2, respectively, and take values in a topological space Y. Let they coincide almost everywhere and the function fl be measurable. Then the function 12 is measurable too. Proof. Let A = {t: t E M2 \ M o, h(t) E U}, Mo = {t: t E Ml n M 2, fl(t) = f2(t)}.
Derivation and integration.
105
The estimate J.t(A) ~ J.t(M2 \ Mo) = J.t(M2 ) - J.t(Mo) = 0 follows from Lemma 1.1 and Theorem 1.1. For every open set U of the space Y we have
Lemma 1.9 implies the measurability of the set f2- 1 (U). The lemma is proved. • Example 2.2. Define on the real line the function X ('Dirichlet function') in the following way. Let
X(t)={~
if t is irrational if t is rational.
The measure of the set of rational numbers is equal to zero. Hence the Dirichlet function coincides almost everywhere with the function identically equal to 1. Lemma 2.1 implies the measurability of the Dirichlet function. Theorem 2.1. Let continuous mappings fl and f2 of a segment [a, b], a < b, into a Hausdorff space Y coincide almost everywhere. Then the mappings fl and f2 coincide. Proof. Lemmas 1.3 and 1.1 and Theorem 1.1 imply that the segment [a, bJ itself is the unique closed subset of full measure of the segment. The assertion in question follows from Lemma 1.7.4. The theorem is proved . • Example 2.3. Let M = [0, IJ U {2}. The functions fl(t) == 1 and
h(t) =
{~
if t E [O,IJ if t = 2
are continuous. They coincide almost everywhere but they are different. Theorem 2.2. Let a function f be defined on a measurable set M ~ lR and take values in a topological space Y with a countable base. Then the following properties are equipotent: a) the function f is measurable; b) there exists a base of the space Y, for every element U of which the set f- 1 (U) is measurable; c) there exists a sub-base of the space Y, for every element U of which the set f- 1 (U) is measurable. Proof. a=? b =? c. This is obvious because every element of a base is an Open set and every base is a subbase. b =?a. Let f3 be a base occurring in condition b, V be an arbitrary open subset of the space Y. Denote by f30 the set of (all) elements of the base f3 lying in the set V. The family f30 covers the subspace V. Theorem 1.6.10 implies the existence of an (at most) countable subfamily f31 of f3o, which
106
CHAPTER 4
also covers the set V. Theorem 1.2 implies the measurability of the set f-l(V) = U{f-l(U): U E ,Bd. c =?b. Let ,B be a sub-base from c. By the definition of a sub-base the set ,Bo of intersections of all finite subfamilies of ,B is a base. Theorem 1.3 implies that for every element V = U1 n ... n Us of the base ,Bo (where U1 , ... ,Us E ,B) the set f-l(V) = f-1(Ud n ... n f-l(Us ) is measurable. The theorem is proved. • Theorem 2.3. Let a set M ~ IR be the union of an (at most) countable family of measurable sets {Mk : k = 1,2, ... }. Let a function f be defined on the set M and take values in a topological space Y. Let the function f be measurable on every set M k , k = 1,2, .... Then the function f zs measurable. Proof. For every open set U of the space Y we have f-l(U)
= U{f-l(U) n M k :
k
= 1,2, ... }.
Theorem 1.2 implies the measurability of the set f-l(U). The theorem is proved. • Corollary. Let a set M ~ IR be the union of an (at most) countable family of measurable sets {Mk: k = 1,2, ... }. Let a function f be defined on the set M and take values in a topological space Y. Let f be constant on every set MkI k = 1,2, .... Then the function f is measurable. • Theorem 2.4. Let measurable functions h, k = 1,2, ... , be defined almost everywhere on a measurable set M ~ IR and take values in a metric space (Y, p). Let the sequence {h: k = 1,2, ... } converge for almost all t E M to a function f (i.e., for almost all t E M we have fk(t) ---+ f(t)). Then the function f is measurable. Proof. The hypotheses and the conclusion of the theorem remain true under the change of the stated functions to functions which coincide with them almost everywhere (see Lemma 2.1). Thus it is sufficient to prove the theorem assuming in addition that the functions are defined on the entire set M and that the needed convergence is at every point of the set M. Let U be an arbitrary open subset of Y. If U = Y then the measurability of the set f-l(U) =M is obvious. In the opposite case by Corollary 2 of Lemma 2.3.1 the function cp(t) = p(t, Y \ U) is continuous. The sets Uk = {t: t E Y, cp(t) > 2- k }, k = 1,2, ... , are open. The sets Hk = {t : t E Y, cp(t) ~ 2- k }, k = 1,2, ... , are closed, U1 ~ HI ~ U2 ~ H2 ~ ... ~ U and U~lUk = U. If x E f-I(U) then f(x) E Urn for some m = 1,2, ... and Urn is a neighborhood of the point f(x). The convergence fk(x) ---+ f(x) implies the existence of a number n ~ 1,2, ... such that for every index k ~ n the point fk(x) belongs to Um. Therefore x E U~=I U~=l n":=nl;;l(Um ). )
Derivation and integration.
107
If x E M \ f- 1 (U) then for every index m = 1,2, ... the set Y \ Hm is a neighborhood of the point f(x). The convergence fk(X) - t f(x) implies the existence of a number n = 1,2, ... such that for every index k ~ n we have h(x) ¢ Hm ;2 Urn. Thus for every n = 1,2, ... the point x does not belong to the set n~nfk1(Urn). Hence x ¢ U~=l U~=l n'k=nfk- 1 (Um ). So f-1(U) = U~=l U~=1 n'k=nfk- 1(Um ). The assertion now follows from Theorems 1.2 and 1.3. The theorem is proved. •
Lemma 2.2. Let Yi and Y2 be topological spaces. Let a set M ~ ~ be measurable. Let a mapping f : M - t Yi be measurable and a mapping g: Y1 - t Y2 be continuous. Then the mapping gf : M - t Y2 is measurable. Proof. Let U be an arbitrary open subset of the space Y2 • By Theorem 1. 7.5 the set g-l (U) is open in Y1. The definition of a measurable mapping implies the measurability of the set (gf)-l(U) = f-1(g-1(U)). The lemma is proved. • Theorem 2.5. Let a function f be defined on a measurable set M ~ ~ and take values in the Euclidean space ~n: f (t) = (I1 (t), ... , f n (t)). The function f is measurable if and only if each function f1, . .. ,fn : M - t ~ is measurable. Proof. Necessity follows from Lemmas 2.1.1 and 2.2. Sufficiency. Let U be an arbitrary open subset of the line, i = 1, ... ,n. We have f- 1 (~x ... x ~ x U x ~ x ... x ~) = fi- 1 (U), where the factor U of the product is in the i-th position. Sets of the form ~ x ... x ~ x U x ~ x ... x ~ constitute a sub-base of the space ~n. Therefore the stated representation of the preimage of an element of this subbase, our assumptions about the functions f1, ... , fm and Theorem 2.2c imply the measurability of the function f. The theorem is proved. • This theorem, Lemma 2.2, and the continuity of operations of the addition and of the multiplication imply the following two assertion. Theorem 2.6. Let a set M ~ ~ and functions f, 9 : M - t ~n be measurable, A E R Then the functions f + 9 and Af are measurable too .• Theorem 2.7. Let a set M ~ ~ and functions f, 9 : M - t ~ be measurable. Then their product f 9 is a measurable function. • These assertions give a large reserve of measurable functions. The continuity of the function h(t) = t- 1 for t :I 0, the continuity of the multiplication, Lemma 2.2 and Theorem 2.5 imply the measurability of the fraction with a measurable function in the numerator and with a measurable nonvanishing function in the denominator. The function h(t) = It I is continuous. Lemma 2.2 implies that if a real function f(t) is measurable then the function If(t)1 is measurable too. Next, if f(t) and g(t) are measurable real functions then the last remark and
108
CHAPTER 4
Theorem 2.6 imply the measurability of the function.
'Pl(t)
= min{J(t),g(t)} =
f(t)
+ g(t) ~If(t) -
g(t)l,
'P2(t)
= max{J(t),g(t)} =
f(t)
+ g(t) + If(t) -
g(t)l.
2 In particular, for every measurable function f(t) its non-negative f+(t) = max{J(t), O} and non-positive f-(t) = min{J(t) , O} parts are measurable too. Let us finish this section by the following remark. Theorem 2.8. Let a real function f be defined on a measurable set M ~ lR. Then the following conditions are equipotent: a) the function f is measurable; b) for every number c E ~ the set {t: t E M, f(t) > c} is measurable; c) for every number c E ~ the set {t: t E M, f(t) ::::; c} is measurable; d) for every number c E ~ the set {t: t E M, f(t) < c} is measurable; e) for every number c E ~ the set {t: t E M, f(t) ~ c} is measurable. Proof. a =*b. This follows immediately from the definition of a measurable function and the openness of the subset (c, 00) of the line. b ¢::}c and d ¢::}e by virtue of the measurability of the complement to a measurable set (see Lemma 1.8). c =* d by virtue of the representation
{t: t E M, f(t) < c}
= U{{t:
t E M, f(t) ::::; c - 2- k }: k
= 1,2, ... }
and Theorem 1.2. e =* b may be proved analogously to the previous implication. Thus conditions b,c,d and e are equipotent. b,c=*a by Theorem 2.2c. The theorem is proved.
•
3. Derivation of non-decreasing functions Let the domain of a real function f contain a set M ~ lR. The function f is called: increasing on the set M if f(s) < f(t) for every two points s < t of the set M; non-decreasing on the set M if f(8) ::::; f(t) for every two points 8 < t of the set M; decreasing on the set M if f(8) > f(t) for every two points s < t of the set M; non-increasing on the set M if f(8) ~ f{t) for every two points 8 < t of the set M.
Derivation and integration.
109
All such functions are called monotone on the set M. Increasing or decreasing functions (on the set M) are also called strongly monotone (on the set M). Let a set M ~ lR lie in the domain of a real function I. A point s of the set M is called a right increasing point (of the function 1 on the set M, respectively, a left increasing point if on the right (respectively, on the left) of the point s there is a point t E M such that I(t) > I(s). If the domain of the function 1 coincides with the set M then usually we omit the mention of the set M in these terms. Lemma 3.1. Let a real function 1 be defined and be continuous on the segment [a, b], a < b, of the real line. Then the set S of right increasing points (respectively, left increasing) is open (in the segment [a, b]). If a < {3 are endpoints of a connected component of the set S then 1 (a) ::;;; 1 ({3) (respectively, 1 (a) ~ f ({3)). Proof. If s is a right increasing point then there exists a point t E (s, b] such that I(t) > f(s). By virtue of the continuity of the function f there exists a number c: > 0 such that f(OES) ~ (-oo,f(t)). Then OES ~ S. This means (in view of the arbitrariness of the point s E S) the openness Figure 4.4 of the set S. Let us prove the last assertion of the lemma. Take an arbitrary point s E (O'.,{3). The number r = sup/([s,b]) is greater than I(s). By Remark 2.3.4 r E I([s, b]). Let t = inf(j-l(r) n [s, b]). Then [s, t) ~ Sand t rf. S. Therefore {3 = t. In view of the arbitrariness of the point s E (a, {3) we have 1((O'.,{3)) ~ (-oo,r). The closed ness of the set (-oo,r] implies the inclusion 1([0'., {3]) ~ (-00, r]. Therefore f(O'.) ::;;; r = f(t) = 1({3)· The function g(x) = f( -x) is defined on the segment [-b, -a]. Left increasing points of the function f and right increasing points of the function g on the segment [-b, -a] are symmetric with respect to zero. Therefore the proved fact implies the assertion of the lemma for left increasing points. The lemma is proved. • Let us go on to the question about the existence of a derivative of a real function 1 defined on a segment of the real line. The derivative at a particular point need not exist. Corresponding examples are well known and we will not discuss them. But if a point x in question is not an endpoint of the segment then in every case we can define the lower )../ and the upper A/ limits of the quotient J(tt:~("') as the point t tends to the point x from the left and the lower )..r and the upper Ar limits of this quotient as the
110
CHAPTER 4
point t tends to the point x from the right, i.e.,
Al = lim inf f (t) - f (x) , t-+x, t<x t - x
f(t) - f(x) · A 1= 1Imsup ,
Ar = lim inf f (t) - f (x) , t-+x, t>x t - x
Ar = limsup f(t) - f(x). t-+x, t>x t- x
t- x
t-+x, t<x
Example 3.1. Consider the functions
f(x) = {xosin
() { M
9 x =
o
~
for x =/: 0, for x = 0,
sin ~ x
=/: 0,
for x
for x = 0,
on the segment [-1,1]. For the function f at the point x = 0 we have Al = Ar = -1, Al = Ar = 1 (Figure 4.5). For the function 9 at the same point we have Al = Ar = -00, Al = Ar = +00 (Figure 4.6).
! .
.
' HI
, "
+- - ; .
Figure 4.5
(, 'I
lil[!:: ; 1[. .~,
~ iV" ': J I! ..
\/
Figure 4.6
For the existence of the (finite) derivative of a function f at a point x the fulfilment of the condition -00 < Al = Al = Ar = Ar < 00 is necessary and sufficient. Lemma 3.2. Let a non-decreasing real function f be defined and be continuous on the segment [a, b], a < b, of the real line. Then the function f has the derivative at almost every point of the segment [a, bJ. Proof. I. Show first that almost everywhere Ar < 00. Let A be the set of (all) points of the segment [a, bJ, at which Ar = 00. For every point x of the set A and every positive integer k there exists a )
Derivation and integration.
111
point t E (x, b] such that
f(t) - f(x) > k. t-x This means that f(t) - kt > f(x) - kx, i.e., the point x is a right increasing point for the function 9k(S) = f(s) - ks. By Lemma 3.1 the set Ak ofright increasing points of the function 9k is open. The point b cannot be right increasing point. We obtain from Corollary of Theorem 2.6.7 and Theorem 2.6.3 that the set Ak may be represented as the union of an (at most) countable set 'Yk of its (pairwise disjoint) components. The components are either intervals or half intervals with the closed endpoint a (no more than one such components). By Lemma 3.1 for every element of the set 'Yk with endpoints a < 13 we have 9k(a) ~ 9k(f3), 13 - a ~ Hf(f3) - f(a)). In view of the monotonicity of the function f this implies that d("(k) ~ tU(b) - f(a)). So J.L*(A) ~ J.L*(Ak) ~ Hf(b)- f(a)). Since this is true for every k = 1,2, ... , J.L*(A) = O. This was required. II. Show that almost everywhere Ar ~ A/. At the points a and b one of these numbers is not defined. Denote by B the set of all points of the interval (a, b), at which the inequality under consideration is not fulfilled, i.e., we have there the inequality A/ < Ar • Denote by P the set of all couples p = (q, r) of rational numbers, in which the first member is strongly less than the second one: q < r. The set P is countable (see Example 1.1.2 and Corollary of Theorem 1.1. 7). The set B may be represented as the union of (countable number of) sets B p , p = (q, r) E P, where Bp is the set of points, at which A/ < q < r < Ar . Our aim will be achieved when we show that the outer measure of every set B p , PEP, is equal to zero. Assume the opposite, i.e., that J.L*(Bp) = C > 0 for some p = (q, r) E P. There exists a covering "I of the set Bp by intervals such that d(,,() < ~c (with the notation of §1). Let V = (a,f3) E "I, By = Bp n V and x E By. At the point x we have A/ < q. Thus there exists a point t E (a, x) such that J(tt~(x) < q and hence f(t) - qt > f(x) - qx, i.e., the point x is a left increasing point of the function f(s) - qs on the segment Iv = [a, 13]. By Lemma 3.1 the set Mv of (all) such points is open in the segment Iv. The left endpoint of the segment Iv cannot be an left increasing point. Corollary of Theorem 2.6.7 and Theorem 2.6.3 imply that the set Mv may be represented as the union of an (at most) countable set "Iv of its (pairwise disjoint) components. The components are either intervals or an half interval with the closed endpoint sup Iv. By Lemma 3.1 for every element U of the set "Iv with endpoints a' < 13' we have the inequality f(a') - qa' ~ f(f3') - qf3' or (3.1)
f(f3') - f(a')
~
q(f3' - a').
CHAPTER 4
112
By virtue of the inequality Ar > r for every point x E UnBir there exists a point t E (x, {3') such that f(tt:~(x) > r, i.e., f(x) - rx < f(t) - rt. Thus the point x is a right increasing point for function f (s) - r s on the segment [a', {3']. By Lemma 3.1 the set M*(c/,{3') of (all) such points is open in the segment [a', {3']. By Corollary of Theorem 2.6.7 and by Theorem 2.6.3 the set M*(a', {3') may be represented as the union of an (at most) countable set ,*(a',{3') of its (pairwise disjoint) components. The components are either intervals or an half interval with the closed endpoint, which coincides with a'. By Lemma 3.1 for every element of the set (a', {3') with endpoints a" < {3" we have the inequality f(a") - ra" ~ f({3") - r{3", i.e.,
,*
{3" - a"
~ ~(f({3") r
The monotonicity of the function
d(r*(a', {3'))
f(a")).
f implies that
~ ~(f({3') r
f(a')).
By (3.1) the total length of elements of the set ,~= U{r*(a',{3'):
(a',{3') E 'v, ( (a',{3] E 'v)}
does not exceed ;J.L(V). The family ,* = U{rir: V E ,} covers the set Bp. By Lemmas 1.2 and 1.1 (and the measurability of open sets)
The contradiction obtained (J.L*(Bp) < J.L*(Bp)) shows that our assumption is false. Hence J.L*(Bp) = O. III. When we go from the function f (x) to the function (x) = - f ( - x ) the notions of left and right change places. From this observation we obtain immediately from the previous reasoning that at almost every point of the segment [a, b] we have Al ~ Ar . When we compare the above inequalities with the obvious estimates Al ~ Al and Ar ~ Ar we obtain that 0 ~ Ar ~ Al ~ Al ~ Ar ~ Ar < 00 at almost every point of the segment [a, b]. So 0 ~ Ar = Al = Al = Ar < 00. Therefore at almost every point of the segment [a, b] the derivative exists. The lemma is proved. • It is natural to ask in what measure the derivative describes the behavior of a function. There exist situations being a long way from the ones which we met when we considered continuously differentiable functions. For the construction of a corresponding example we will need
r
Derivation and integration.
113
Lemma 3.3. Let M be a closed subset of the segment [a, b], a < b. Let a real function f be defined and be non-decreasing on the segment [a, b], be continuous on the set M and on the closure of every component of the set [a, b] \ M. Then the function f is continuous. Proof. The continuity of the function f at points of the set [a, b] \ M follows from the openness of the set [a, b] \ M in the segment [a, b] and hypotheses of the lemma. Thus we need to prove the continuity of the function f at an arbitrary point x of the set M. Take an arbitrary number c > O. By virtue of the continuity of the function f on the set M there exists a number 8 > 0 such that If(t) - f(x)1 < c for every t E (x - 8, x + 8) n M. If (x - 8, x) n M "# 0 then denote by c an arbitrary point of the set (x - 8, x) n M. If (x - 8, x) n M = 0 then by virtue of the continuity of the function f on the closure of the component of the set [a, b] \ M, which contains the interval (x - 8, x), there exists a number c E (x - 8, x) such that If(c) - f(x)1 < c. Thus in this case too we have fixed a point c. If (x, x + 8) n M "# 0 then denote by d an arbitrary point of the set (x,x + 8) n M. If (x,x + 8) n M = 0, then by virtue of the continuity of the function f on the closure of the component of the set [a, b] \ M, which contains the interval (x, x + 8), there exists a number d E (x, x + 8) such that If(d) - f(x)1 < c. Thus in this case too we have fixed a point d. The interval (c, d) is a neighborhood of the point x. By virtue of the monotonicity of the function f we have
f(( c, d))
~
[f( c), f(d)]
~
(f(x) - c, f(x)
+ c).
The lemma is proved. • Example 3.2. We will construct a continuous mapping of the segment [0, 1] onto itself, a so called 'Cantor stairs'. Use the notation of Examples 1.6.4 and 2.2.8. In discussion of properties of this function the result of Example 1.2 will be also essential: the measure of the Cantor perfect set is equal to zero. In Example 2.2.8 we have pointed an homeomorphism f : K --t DN. Define first the function h ('Cantor stairs') on the set K and next extend it to the entire segment [0, 1]. For a point x = {Xl, X2, ... } E DN put Sets VX1, ... ,Xk
=
{y
= {YI, Y2,'" }: Y E D N , YI = Xl,··· ,Yk =
= {Yl, Y2,' .. } E VX1, ... ,Xk' then h ( )1 ~ IXk+l - Yk+d + IXk+2 - Yk+21 + 0 Y '" 2k+1 2 + 2 ' ..
constitute a base at the point x. If Y
Ih o( x ) -
xd
k
~
111 + 2k+2 + ... = 2k '
2k+l
CHAPTER 4
114
This implies the continuity of the mapping ho and hence the continuity of the mapping h = hof. Let [e, d] be a component of the set Sk([O, 1]). Then (with the notation of Example 2.2.8) for every i = 1, ... , k the set fi([e, d]) consists of one point. Denote it by ai. The values of the mapping f at the points e and d may be written as
f(e) = {al, ... ,ak,O,O, ... }, Let el
f(d) = {al, ... , ab1, I, ... }.
= e + ~(d - e), d l = e + Hd - e). We have:
f(el) = {al, ... ,ak,O,I,I, ... }, Therefore
and (3.2)
h() = h(d ) = al el
1
2
+
...
ak
_1_ = h(e)
+ 2k + 2k+l
+ h(d) . 2
Thus values of the function h at the endpoints of a component (el, dd of the complement to the Cantor perfect set coincide. Extend the function h to the segment [0,1] and put h(x) = h(ed for every x E (el,dd. The equality (3.2) helps to imagine the graph of this function (Figure 4.7). The function h is monotone. Lemma 3.3 implies its continuity. The continuity, the equalities h(O) = and h(l) = 1, Corollary 1 of TheoFigure 4.7 rem 2.6.4, and Theorem 2.6.3 imply the coincidence h([O, 1]) = [0,1]. The complement of the Cantor perfect set in the segment [0,1] has the full measure. At every its point the derivative of the function h exists and is equal to zero. We have constructed the continuous non-decreasing nonconstant function, the derivative of which almost everywhere is equal to zero. Recall for comparison that if the derivative of a function exists and is equal to zero everywhere (but not 'almost everywhere') then the function
°
Derivation and integration.
115
is constant. It seems that Lemma 3.2 gives a possibility to introduce the notions of the integral and of the primitive for a new rather large class of functions. Example 3.2 shows that here we go away from relations between functions and derivatives to which we are accustomed in dealing with continuously differentiable functions. However in Lemma 3.2 we made a step in the direction needed. Our nearest aim will be to select a class of functions which have derivatives almost everywhere and which are uniquely determined by their derivatives up to constant terms. We will pass to it in the next section, but now Theorem 3.1. A convergent series of continuous non-decreasing functions f(x) = fl (x) + f2(X) + h(x) + ... on a segment [a, b], a < b, admits almost everywhere the termwise derivation:
f'(x)
= f~(x) + f~(x) + f~(x) + ....
Before to prove the theorem let us make some introductional remarks. As usual, the sum of the series means the limit of the sequence of the partial sums
as k -+ 00. In the statement of the theorem we did not define exactly the type of convergence of the series. Assume now the convergence of the series at points a and b only. By virtue of the monotonicity of the functions fk' k = 1,2, ... for every point t of the segment [a, b] and for every index i < j we have In view of the arbitrariness of the point t E [a, b] this implies that
where in the left hand side the norm of the uniform convergence appears. This inequality, the convergence of the sequences {s d a): k = 1, 2, ... } and {s k (b): k = 1, 2, ... }, and the fact that every convergent sequence is fundamental (see §1.6) imply that the sequence {Sk: k = 1,2, ... } is fundamental with respect to the norm of the uniform convergence. From Corollary of Theorem 3.1.2 we obtain the convergence of the sequence {Sk : k = 1,2, ... } to a continuous function f. This guarantees the summability of Our series with respect to the norm of the uniform convergence. The sum f of our series is non-decreasing. By Lemma 3.2 terms of the series and the sum have derivatives almost everywhere on the segment [a, b]. Proof of Theorem 3.1. Consider a subset M of full measure of the segment [a, bj, at points of which derivatives of all terms and of the sum of
CHAPTER 4
116
the series exist. For every point x E M and for every point t E [a, b] \ {x} we have =
t-x
t- x
Signs of terms of the numerator in the left hand side coincide with the sign of the denumerator. Thus for every number N
When we pass to the limit as t
--+
x we obtain
N
L f~(x) ~ J'(x). k=1
Since this is true for every N = 1,2, ... the nonnegativity of the terms implies the summability of the series of the derivatives and the estimate (3.3)
f{(x) +
f~(x)
+ f~(x) + ...
~
J'(x).
Now for every k = 1,2, ... fix a number i(k) = 1,2, ... such that SiCk) II ~ 2- k. The above remarks may be applied to the series
Ilf -
g(x) = gl (x) + g2(X) + g3(X) + ... , where gk(X) = f(x) - Si(k)(X) = fi(k)+l(X) + fi(k)+2(X) + .... Thus at almost every point x of the segment [a, b] we have
g'(x)
~ g~ (x)
+ g~(x) + g;(x) + ...
The convergence of a numerical series implies that their terms tend to zero. In our case this means that S~(k)(X) --+ f'(x) as k --+ 00. Now (3.3) and the nonnegativity of terms of the series in the left hand side of the inequality (3.3) imply that Si(X) --+ f'(x) as i --+ 00 for almost all x E [a, b]. The theorem is proved. • 4. Derivative with respect to a set Let the domain of a function f with values in the Euclidean space ~n (for n = 1 in the real line ~) contains a set M ~ R Let x be a non-isolated point of the set M. If the limit of the quotient I(tt:~("') as t E M, t --+ x exists, it is called a derivative of the function f with respect to the set M at the point x and is denoted by f~(x).
117
Derivation and integration.
Following the standard method we prove that if two functions have the derivatives with respect to a set M ~ IR at a point x E M then their sum, difference, and product of each of these functions by a number also have such a derivative and its value is equal, respectively, to the sum, the difference and the product by the number of the values of the derivatives of initial functions. A mapping f (t) = {fl (x), ... , f n (t)} of the set M into the space IR n has the derivative at a point x E M with respect to the set M if and only if the coordinate functions fl(X), ... , fn(x) have the corresponding derivatives (necessity follows from Lemmas 2.1.1 and 1.7.2, sufficiency follows from Lemma 2.1.2). In this case fk(x) = {(fl)~(X), ... , (fn)~(x)). Lemma 4.1. Let the domain of a function f with values in the Euclidean space IR n contain sets A ~ B (~ 1R). Let x be a non-isolated point of the set A (hence x is a non-isolated point of the set B too). Let the derivative f~(x) exist. Then the derivative f~(x) exists and f~(x) = f~(x). Proof. The proof is obvious. • Lemma 4.2. Let the domain of a real function f contain a (nonempty) compact set M ~ R Let the function f be continuous and non-decreasing on the set M. Then the function f has a derivative with respect to the set M at almost every point of M. Proof. It is sufficient to prove the assertion under the additional assumption 7r(f) = M. Let a = inf M and b = supM. The set L = [a, bj \ M is open in the segment [a, bj and does not contain the points a and b. Theorems 2.6.7 and 2.6.3 imply the representation of the set L as the union of a set 'Y of pairwise disjoint intervals. Endpoints of these intervals belong to M. Extend the function f to the entire segment [a, bj by putting f(t) = f(a)({3 - t) + f(f3)(t - a) (3-a
for t E [a,bj\M, where t E (a,{3) E 'Y, see Figure 4.8. By Lemma 3.3 the extended function f is continuous. Evidently it is non-decreasing. By Lemma 3.2 there exists a set Mo ~ [a, bj of measure zero such that for every point x of the set [a, bj \ Mo the derivative f'(x) exists. Lemma 4.1 implies the existence of the derivative fk(x) for every point x E M \ Mo. It remains to notice, that by Theorem 1.1 J.t(M \ Mo) = J.t(M). The lemma is proved. •
-M a
M
fJ Figure 4.8
M
CHAPTER 4
118
Lemma 4.3. Let the domain of a function f with values in the Euclidean space Rn contain a measurable set M ~ R Let the (finite) derivative f~ (x) exist at almost every point x of the set M. Then the function f~ is measurable. Proof. Let Mo be the set of all points x EM, at which the derivative f~(x) exists. If x E M o, then the quotient f(tt:~(x) tends to a finite limit, f(t) - f(x) = f(t) - f(x) (t - x) t-x
~
0
as t ~ x in the set M. Therefore the function f IM is continuous at x. Fix a point tik E (2- i k, 2- i (k+1))nM for i = 1,2, ... , k = 0, ±1, ±2, ... satisfying the condition (2- i k, 2- i (k + 1)) n M =I 0. The set Ml of all so fixed points is (at most) countable. Therefore IL(Md = O. For i = 1,2, ... , k = 0, ±1, ±2, ... and x E (2- i k, 2-i(k + 1)) n (Mo \ (Ml U Q)) put gi(X) = f(x) - f(t ik ). x - tik
The function gi is continuous on every set
i = 1,2, .... Therefore it is continuous (see Theorem 1.7.7). Thus the func-
tion is measurable (on Mo \ (Ml U Q)). For a fixed x E Mo \ (Ml U Q) by virtue of the existence of the derivative f~(x) the sequence gi(X) converges to f~(x). Theorem 2.4 implies the measurability of the function f~IMo\(MluQr Lemma 2.1 implies the measurability of the function f~· The lemma is proved. • 5. Absolutely continuous functions Let the domain of a function f with values in a metric space (X, p) contain a set M ~ R The function f is called absolutely continuous on the set M if for every number c > 0 there exists a number > 0 such that if
°
(*,0, M) a finite family,
= {(ai, (3i):
i
= 1, ... ,k} consists of pairwise
disjoint intervals with endpoints from the set M and with d(,) < 0, then E{p(f(ai),f({3;)): i
= 1, ... ,k}
~
c.
If the set M coincides with the domain of the function f then the function f is called absolutely continuous (without the mentioning of M). Lemma 5.1. Let a function f with values in a metric space (X, pd be absolutely continuous on a set M ~ R Let a mapp'ing 9 of the space X into
Derivation and integration.
119
a metric space (Y, P2) satisfy the Lipschitz condition. Then the function gf is absolutely continuous on M. Proof. Take an arbitrary number c > O. Let L (~ 0) be a Lipschitz constant for the mapping g. Select a number 0 > 0 such that as soon as a family'Y = {(ai, .Bi): i = 1, ... ,k} satisfies condition (*,0, M), then
For every such a family 'Y 'L3p2(gf(ai), gf(.Bi)): i = 1, ... ,k}
~ L 'L3P1 (f(ai), f(.Bi)): i = 1, ... , k}
~ L
Lc
+ 1 < c.
The lemma is proved. • Lemma 5.2. The diagonal product f of functions fj, j = 1, ... ,n, with values in metric spaces (Xj, pj) is absolutely continuous on a set M ~ IR (with respect to the metric on the product from Theorem 2.1.2) if and only if every function Ii is absolutely continuous on M. Proof. Necessity follows immediately from Lemma 5.1. Sufficiency. Let P be the metric on the product. Take an arbitrary number c > O. For every j = 1, ... ,n fix a number OJ > 0 in a manner that
for every family'Y = {(ai,.Bd: i = 1, ... , n} satisfying condition (*, OJ, M). If 0 = min{ 01,'" ,On} and a family 'Y = {(ai, .Bi): i = 1, ... ,k} satisfies condition (*,0, M), then L{p(f(ai),f(.Bi)): i = 1, ... ,k}
~L
{t
pj(Jj(ai), Ii (.Bi)) : i
= 1, ...
'k}
3=1
n
= LL{pj(Jj(ai),fj(.Bi)):
i
= 1, ... ,k}
i=1 ~
c n· n
The lemma is proved.
= c.
•
120
CHAPTER 4
Lemma 5.3. A vector function f = {fl, ... , fn} with values in the Euclidean space IRn is absolutely continuous on a set M ~ IR if and only if every coordinate function fl, ... , fn is absolutely continuous on M. Proof. To be precise, in the statement we speak about the Euclidean metric of the space IRn. For the metric of Example 1.4.5 an analogous assertion follows immediately from Lemma 5.2. Because of the inequality (2.2.1) relating these two metrics the assertion in question follows from Lemma • 5.1. The lemma is proved. Lemma 5.4. Let the domains of functions f and 9 contain a set M ~ R Let the functions take values in the Euclidean space IRn and be absolutely continuous on the set M. Let>. be a real number. Then the functions f + 9 and >.f are absolutely continuous on the set M. Proof. Lemma 5.3 reduces the proof to the scalar case. Here we obtain what was required from Lemma 5.2, remarks of Examples 1.7.2 and 1.7.3 and Lemma 5.1. The lemma is proved. • Example 5.1. The identity mapping of a set M ~ IR into itself is absolutely continuous. This follows immediately from definitions. So Lemma 5.1 implies that if a function satisfies the Lipschitz condition then it is absolutely continuous. Assume that a function f is defined, has the continuous derivative on the segment [a, bj, and m = sup f'([a, b]). For every point s < t of the segment [a, bj there exists a point c E [s, tj such that f(t) - f(s) = f'(c)(t - s) (Lagrange formula). So If(t) - f(s)1 ~ mit - 81, i.e., the function f satisfies the Lipschitz condition. Hence it is absolutely continuous. Lemmas 5.4 and 5.3 extend immediately our list of examples of absolutely continuous functions. Every absolutely continuous function is uniformly continuous. We obtain it, when we take in (*,0, M) a family 'Y consisting of one element. Lemma 5.5. Let a function f with values in a metric space (X, p) be absolutely continuous on a set M ~ R Then for every number EO > 0 there exists a number > 0 such that if
°
(**,o,M) an (at most) countable family 'Y = {(ai,,Bi): i = 1,2, ... } consists of pairwise disjoint intervals with endpoints from the set M and with d(,) < 8, then 'L-{p(f(ai), f(,Bi)): i = 1,2, ... } ~ EO. Proof. For an arbitrary number EO > 0 find a number 8 > 0 according to the definition of the absolute continuity. If a family 'Y = {( ai, ,Bi) : i = 1,2, ... } is infinite and satisfies condition (**,8, M) then for every k = 1,2, .. .
Derivation and integration. When we pass to the limit as k
-t
00,
121
we obtain
(For a finite family, we obtain what was required directly from the definition.) The lemma is proved. • Theorem 5.1. Let a real function j be defined and be absolutely continuous on a segment [a, b]' a < b, of the real line. Then for every number c > 0, there exists a number 8 > 0 such that if M ~ [a, b] and J.L*(M) < 8, then J.L*(f(M)) :::;; c. Proof. I. For arbitrary c > 0 fix 8 according to Lemma 5.5. Let M ~ [a, b] and J.L*(M) < 8. II. Consider the particular case, where the set M is open in the segment [a, b]. By Corollary of Theorem 2.6.7 the set M may be represented as the union of an at most countable set, of its connected components. By Theorem 2.6.3 every element J of the set, is either an interval or a half interval or a segment. Let a(J) and (3(J) be its initial point and the end, s(J) = inf j([a(J), (3(J)]) and t(J) = sup j([a(J), (3(J)]) , see Figure 4.9. Remark 2.3.4 By s(J) = j(a'(J)) and t(J) t( J) = j ({3' (J)) for some cl(J), (3' (J) E [a( J), (3(J)] 1(J) (Figure 4.9). This means that under the action of j the image of the segs(J) ment I(J) with the endW(J) ~ a.'(J) . points a' (J) and (3' (J) co---------------a.(J)-;---~(J) incides with the image of the segment [a(J),{3(J)]. Figure 4.9 By the choice of 8 and by results of §1
J.L(f(M)) :::;; J.L(U{f([a(J),{3(J)]): J E ,}) = J.L(U{[s(J), t(J)]: J E ,}) :::;; 2)If(a'(J)) - j({3'(J))I: J E ,} :::;; c, because I(J) ~ [a(J), (3(J)] and
I)Ia'(J) - {3'(J)I: J E ,} :::;; 2)la(J) - {3(J)I: J E ,}
= J.L(M) < 8.
III. Let us go on to the general case. The definition of the outer measure implies the existence of an at most countable family of intervals, covering the set M with db) < o. By Theorem 1.2 the measure of the set G 1 =
U,
CHAPTER 4
122
is less than o. The measure of the (see Lemmas 1.7 and 1.1). The set definition of the induced topology. theorem is proved. Corollary. Let a real function f
set G 2 = G 1 n [a, bj is less than 0 too G 2 is open in the segment [a, bj by the By II JL(f(M)) ~ JL(f(G 2 )) ~ E. The •
be defined and be absolutely continuous on a segment [a, b], a < b, of the real line and M be a subset of the measure zero of the segment [a, bj. Then JL(f(M)) = o. • Lemma 5.6. Let a function f with values in a metric space (X,p) be defined and be continuous on a subset M of the real line R Let it be absolutely continuous on a set Mo ~ M. Then function f is absolutely continuous on the set Ml = [Moj n M. Proof. Take an arbitrary number E > O. According to the definition of the absolute continuity find 0 > O. Let a family 'Y = {(ai, ,8i) : i = 1, ... , k} satisfy condition (*,0, Md and s = db) « 0). For every n = 1,2, ... and i = 1, ... ,k fix points ar E O{ai n Mo and ,8': E 0e,8i n M o, where ~ = 2-nk(0 - s), such that intervals of the family 'Yn = {( ar,,8,:): i = 1, ... , k} are pairwise disjoint. We have
The family 'Yn satisfies condition (*,0, M). Thus z)p(f(a~),f(,8~)): i = 1, ... ,k} ~
E.
Since
- p(f(ai), f(,8i)) I ~Ip(f(a~), f(,8~)) - p(f(a~), f(,8i)) I + Ip(f(an, f(,8i)) - p(f(ai), f(,8d) I ~p(f(,8~), f(,8i)) + p(f(a~), f(ai))
Ip(f(a~), f(,8~))
(the last passage is made by Lemma 2.3.1), z)lp(f(a~), f(,8~)) - p(f(ai), f(,8i)) I : i = 1, ... , k} ~ L {p(f(a~), f(ai)): i = 1, ... ,k}
+ L{p(f(,8~),f(,8d): Terms in the right hand side tend to zero as n
L
-+ 00.
Therefore
= 1, ... ,k} {p(f(ai), f(,8i)): i = 1, ... ,k}.
L)p(f(ar), f(,8':)): i -+
i = 1, ... ,k}.
123
Derivation and integration.
When we pass here to the limit, we obtain what was required:
z)p(f(ad,f(,8i)): i = 1, ... ,k}:S; c. The lemma is proved. • Lemma 5.7. Let a function f with values in a metric space X be absolutely continuous on a set M ~ JR. and MI ~ M. Then the function f is absolutely continuous on the set MI' Proof. The proof is obvious. •
6. Derivation of absolutely continuous functions Lemma 6.1. Let a real function f be defined and be absolutely continuous on a (nonempty) compact subset M of the real line JR.. Then there are absolutely continuous and non-decreasing functions g and h defined on the set M such that f = g - h. Proof. 1. For points a < ,8 of the set M denote by V(a,,8) the upper bound of sums
where a = Xo :S; ... :S; Xk+l = ,8 and {Xl,'" , xd ~ M. For now we do not assert that this upper bound is finite. II. Let a :S; ,8 :S; , be points of the set M. Consider an arbitrary sum
where a = xo:S; ... :S; Xk+l =, and {xI, ... ,xd such that X j :S; ,8 :S; X j+l' Then
s:S; (l)lf(xi+l) - f(xi)l: i = O, ... ,j - I}
+ (If(xj+d - f(,8)1 + 2: {If(XHI) :S; V(a,,B) + V(,8,,).
~
M. Fixj
+ If(,8)
= O, ... ,k
- f(xj)l)
f(xi)l: i = j
+ 1, ... , k})
When we pass to the upper bound in the left hand side of this inequality we obtain (6.1)
V(a,,) :S; V(a,,B)
On the other hand, for every c >
V(,8,,) we can fix points a
= Xo
+ V(,8, ,).
°
by the definition of V(a,,8) and
~ ... :S; Xj+l
= ,8
~ Xj+2 ~ ... ~ Xk+l
of the set M such that
L {If(xHd L {If(xHd -
f(xi)l: i = 0, ... ,j} ~ V(a,,8) - c,
!(xi)l: i = j + 1, ... , k} ~ V(,8,,) - c.
=,
CHAPTER 4
124 Therefore
By the definition of V(a,,)
V(a,,)
~
V(a,.8)
+ V(.8,,)
- 2£.
In view of the arbitrariness in the choice of £ > 0 (6.2)
V(a,,)
~
V(a,.B)
+ V(.8, ,).
When we compare (6.1) and (6.2), we obtain
V(a,,) = V(a,.8)
+ V(.8, ,).
III. For an arbitrary number £ > 0 take a number Ii > 0 according to the condition which occurs in the definition of the absolute continuity. Show that as soon as a family, = {( ai, .8i): i = 1, ... , k} satisfies the condition (*, Ii, M) of §5, then
Assume the opposite, i.e., assume the existence of a family , = {( ai,.8d: i = 1, ... ,k} satisfying condition (*, Ii, M) such that the number 'fJ = L {V(ai' .8i) : i = 1, ... , k} - c is positive. By the definition of V(ai' .8i) for every i = 1, ... ,k we can choose points
of the set M in a manner that
The family U{{(Xi,/,Xi,l+d: (*, Ii, M) and the estimate
1 = O, ... ,j(i)}:
~ {If(Xi,I+l) - f(Xi,/)I: 1 = 0, ... ,j(i), i
i = 1, ... ,k} satisfies
=
1, ... , k} > c.
This contradicts the choice of Ii. Thus our assumption is false. This gives what was required. IV. Let a = inf M. For t E M put g(t) = V(a, t). It follows from II and from the nonnegativity of values of V(a, .8), that the function 9 is non-decreasing. Show that for every t E M the value g(t) is finite. The set )
Derivation and integration.
125
Mo = {t: t E M, g(t) < oo} is nonempty because it contains the point a: g(a) = O. Let to = sup Mo.
By virtue of the monotonicity of the function g we have Mn[a, to) ~ Mo. For c = 1 find a number 6 > 0 according to III. Next, fix an arbitrary point s E (to-~, to+~)nMo. For every point t E (s, to+~)nM (and, in particular, for t = to) g(t) = g(s) + V(s, t) ~ g(s) + 1 < 00. Therefore t E Mo. The set Mo is open in the subspace M. If to =1= sup M, then the set M \ Mo is nonempty. In this case by virtue of the openness of the set Mo the point tl = inf(M \ Mo) is different from the point to and (to, td n M = 0. The last equality implies that V(to, td = If(td - f(to)l. So
Hence tl E Mo. This contradicts the definition of the point t 1 . Thus our assumption is false and to = sup M. This means the finiteness of values of the function g. V. The absolute continuity of the function g follows from II and III. VI. The absolute continuity of the functions f and g implies the absolute continuity of the function h = g - f. If points a < (3 belong to the set M, then the definition V( a, (3) implies immediately that If((3) - f(a)1 ~ V(a, (3).
Therefore h((3) - h(a) = (g((3) - g(a)) - (f((3) - f(a)) ~
V(a, (3) - If((3) - f(a)1 ~ 0,
i.e., the function h is non-decreasing. The lemma is proved. • Theorem 6.1. Let a function f with values in the space IR n be defined and be absolutely continuous on a compact set M ~ R Then almost everywhere on the set M the function f has the derivative.
Proof. Lemma 5.2 reduces the proof to the scalar case. Here the assertion follows from Lemmas 6.1 and 4.2. The theorem is proved. • 7. Lebesgue integral Lemma 7.1. Let a real function f be defined and be continuous on a segment [a, b], a < b, of the real line. Let it be absolutely continuous on a closed set M ~ [a, bj and at almost every point of the set M its derivative with respect to the set M be positive. Let the function f be increasing on every component of the set [a, bj \ M. Then the function f is non-decreasing. Proof. If a point x E M \ {b} is not the left endpoint of a component of the complement to the set M (but almost every point of the set M satisfies
126
CHAPTER 4
this condition), the derivative Jk(x) exists and is positive, then there exists a point t E M n (x, b) such that
J(t) - J(x) > t-x
:........:...-'-----=--..:........:...
o.
Hence J(t) - J(x) > 0, i.e., the point x is a right increasing point of the function J. Every point of the complement to the set M (except the point b, if it belongs to the complement) is also a right increasing point. This follows immediately from the hypotheses of the lemma. Thus the set 8 of right increasing points of the function J has the full measure on the segment [a, bj. By Lemma 3.1 the set 8 is open. For endpoints Q < f3 of every its component we have J(Q) ~ J(f3). Enumerate elements of the set 'Y of connected components of the set 8 and for k = 1,2, ... denote by 'Yk the set of components with numbers 1, ... ,k. Elements of the set 'Yk are pairwise disjoint. Therefore their positions in the segment [a, bj are ordered: if we take two its different elements then one of them lies entirely on the left of the another element. Enumerate anew elements of the set 'Yk in order to guarantee that every element with smaller number be placed in the left of every element with larger number: 'Yk = {II, ... ,ld. Introduce the numbers
+ ... + (J(f3~) - J(Q~)), J(f3f)) + ... + (J(Q~) - J(f3LI))'
Pk = (J(f3f) - J(Q~)) qk = (J(Q~) where Q~
< f3; are endpoints of Ii. Since J.t(U'Yk) ~ f3~ - Q~ ~ b - a b- a
f3t -
Q~
---t
= 1'(8) = J.t(U'Y) = "{J.t(l): L....J
and
I E 'Y}
= k--+oo lim J.t(U'Yk),
b - a. The inequalities
a ~ ... ~ Q~ ~ Q~ ~ f3i ~ f3~ ~ ... ~ b, imply that Q~ ---t a and f3t ---t b. Since Pk + qk = J(f3t) - J(Q~), we obtain the convergence Pk + qk ---t J(b) - J(a) as k ---t 00. The endpoints of the intervals (f3f, Q~), ... ,(f3LI' Q~) belong to the set M. Their total length (i.e., by virtue of Theorem 1.1, the measure of their union) is ~ (b - a) - J.t(u'Yd. Since the set 8 has the full measure on the segment [a, b] the total length tends to zero as k ---t 00. Now the absolute continuity of the function J on the set M implies that
Iqkl ~ If(Q~) - f(f3f)1
+ ... + If(Q~)
as k ---t 00. On the other hand, Pk f(b) - f(a) ~ 0, i.e., f(b) ~ f(a).
~
o.
- f(f3:-I)1
---t
0
When we compare, we obtain
Derivation and integration.
127
If we take two arbitrary points 8 < t of the segment [a, b]' then we see that the hypotheses of the lemma hold for the segment [8, t], the function fl[s,t] and the set M n [8, tj. By the just proved assertion f(t) ~ f(8). The lemma is proved. • We can pass immediately to a more precise result. Lemma 7.2. Let a real function f be defined and be continuous on the segment [a, b], a < b, of the real line, be absolutely continuous on the closed set M ~ [a, bj. Let at almost every point of the set M its derivative with respect to the set M be non-negative. Let the function f be non-decreasing on every component of the set [a, b] \ M. Then the function f is non-decreasing. Proof. For every e > 0 the function f(x) + eX satisfies the hypotheses of the previous lemma. Apply it. When we pass to the limit as c ~ 0 we obtain what was required. The lemma is proved. • Corollary. Let a real function be defined and be absolutely continuous on the segment [a, bj. Let at almost every point of the segment [a, bj its derivative be non-negative. Then the function f is non-decreasing. • The following version of Lemma 7.2 will be helpful too. Proposition 7.1. Let a real function f be defined and be absolutely continuous on a compact subset M ~ [a, bj of the real line. Let at almost every point of the set M its derivative with respect to the set M be non-negative. Let f(a) ~ f(fJ) for every component (a, f3) of the set [inf M, sup Mj \ M. Then the function f is non-decreasing on the set M. Proof. Extend the function f on the segment [inf M, sup Mj and define it at a point t of a component (a, fJ) of the set [inf M, sup Mj \ M by the formula
f(t)
=
t-a fJ-t fJ _ a f (f3) - fJ _ af(a).
The so extended function f satisfies the hypotheses of Lemma 7.2 (see • Lemma 3.3). This gives what was required. Lemma 7.3. Let a function f with values in the Euclidean space ~n be defined and be absolutely continuous on a segment [a, b], a < b, of the real line. Let at almost every point of this segment its derivative be equal to zero. Then the function f is constant . Proof. In the scalar case Corollary of Lemma 7.2 implies that both functions f and - f are non-decreasing. This may be only when the function f is constant. In the vector case the hypotheses of the lemma imply that almost everywhere derivatives of the coordinate functions are equal to zero. The previous remark implies that these functions are constant. Thus the function f is constant too. The lemma is proved. •
128
CHAPTER 4
Example 3.2 ('Cantor stairs') shows that in Lemma 7.3 we cannot omit the requirement of the absolute continuity of the function f without its compensation by any other restrictions. Let a function f be defined almost everywhere on the segment [a, b]' a < b, of the real line and take values in the space IRn. The function f is called Lebesgue integrable if there exists an absolutely continuous function F : [a, b] - t IRn such that F'(t) = f(t) for almost every t E [a, b]. The function F is called a primitive (or an indefinite integraQ of the function f. As in the case of the Riemmann integral (Le., of the integral of a continuous function whose theory is usually accounted at first steps of the course of Mathematical Analysis) the primitive is defined up to an additive constant: if F and G are two primitives of a function f on [a, b], then for almost all x E [a, b] we have (F(x) - G(x))' = f(x) - f(x) = O. By Lemmas 7.3 the difference F - G is constant. On the other hand, if F is a primitive of f on [a,b] and c E IRn , then (F(x) + c)' = F'(x) at every x E [a,b]' at which the derivative F'(x) exists. Thus F + c is also a primitive to f. By virtue of the last remark for every points s, t of the segment [a, b] the difference F(t) - F(8) does not depend on the choice of the primitive F. This vector (number in the scalar case) is called a (definite) integral of the function f from 8 to t (or over the segment [8, t], when 8 ~ t). It is denoted by t
J
f(x)dx.
8
Evidently t
J
J
8
t
f(x)dx = -
8
f(x)dx.
The derivation is made in each coordinate independently. Thus the finding of a primitive, which is a vector function, and calculation of a definite integral, which is a vector, may be also made in each coordinate independently. If a function f is integrable on a segment [a, b], then it is integrable on every smaller segment I, i.e., the function fll is integrable. If F be a primitive of the function f then the function FII is a primitive of the function fl/' On the other hand, if 8 E [a, bJ and a function f is integrable on the segments [a,8J and [8, bJ then it is integrable on the segment [a, bJ. Proof of this assertion may be proposed to the reader as an easy exercise. J
129
Derivation and integration.
f
If s, t and u are three arbitrary points of the segment [a, b], a function is defined and Lebesgue integrable on [a, bj, then u
Jf(x)dx = F(u) - F(s) 8
= (F(t) - F(s)) t
=
+ (F(u)
- F(t»
u
J
f(x)dx
+
J
f(x)dx,
t
8
where F is an arbitrary primitive of the function f. Let F be a primitive of a function f on a segment [a, bj and s E [a, bj. For t E [a, bj put t
G(t) =
J
f(x)dx = F(t) - F(s).
s
The function G differs from the primitive F by a constant -F(s). Because of above remarks it is a primitive for the function f on the segment [a, bj (moreover G(s) = 0), i.e., a definite integral when we consider it as a function on its upper limit of integration (i.e., of the number t in the above formula), is a primitive. Let a function f be defined on a interval or on a half interval J. The function f a is called locally integrable if it is integrable on every segment lying in J. For every s E J the formula t
G(t) =
Jf(x)dx s
defines a primitive of the function f on J (i.e., for every segment I ~ J the function Gil is a primitive of the function fII). For a function f being locally integrable on a half interval [a, (0) the limit s
8li.~
Jf(x)dx a
(under the natural assumption that it exists) is denoted by
J 00
f(x)dx.
a
130
CHAPTER 4
Analogously we define the value of b
J
f(x) dx
-00
Put also
a
J
J
00
f(x)dx =
-00
J 00
f(x)dx
+
f(x)dx,
a
-00
where a E (-00,00) (evidently the sum in the right hand side does not depend on the choice of a). In these formulae we can use every other suitable symbol instead of x to denote of the independent variable. Let M be a subset of the set X. The function
X(X)={~
if x E M, if x E X \ M.
is called a characteristic function of the set M (in X). Let M be a (measurable) subset of the real line. We regard the symbol
J
f(x)dx
M
as equipotent to the symbol 00
J
f(x)X(x)dx,
-00
where X is a characteristic function of the set M (in 1R). Example 7.1. Every continuous function is Lebesgue integrable (see Example 5.1). Its Lebesgue integral coincides with its Riemann integral. Example 7.2. Let -00 < a < (3 < 00. The characteristic function of a segment, of an interval or of a half interval with endpoints a, (3 is Lebesgue integrable: the function
F(x) = {x ~ a (3-a
if x < a, if a ~ x ~ (3, if x > (3
is its primitive (the absolute continuity of the function F may be easy checked directly).
131
Derivation and integration.
Theorem 7.1. Let functions f and g with values in the space ~n be defined almost everywhere on a segment [a, b], a < b, of the real line. Let them be Lebesgue integrable. Let>. be an arbitrary real number. Then the functions f + g and >.g (defined almost everywhere on the segment [a, bj) are Lebesgue integrable and b
b
+ g(x))dx
jU(x)
=
a
b
j f(x)dx a
+j
g(x)dx,
a
b
b
j >.f(x)dx = >. j f(x)dx. a
a
Proof. The integrability follows from Lemma 5.4. The equality may be • easily checked directly. The theorem is proved. Lemma 7.4. Let a non-negative function f be defined (almost everywhere) and be integrable on a segment [a, b], a < b, of the real line. Then b
j f(x)dx
~ O.
a
Proof. By Corollary of Lemma 7.2 (each) primitive of the function f is a non-decreasing function, that immediately gives what was required. The lemma is proved. • Theorem 7.2. Let non-negative real functions
j(
+ ... +
a
for every k
= 1,2, ... ). Then for almost all t
converges, the function
f
=
L f
boob
a
E [a, bj the series
k=l
a
CHAPTER 4
132
Proof. By Corollary of Lemma 7.2 each function t
Fk(t) =
J
'h(x)dx
a
is non-decreasing. By remarks preceding the proof of Theorem 3.1 the series FI + F2 + F3 + ... converges uniformly. Theorem 3.1 will imply our assertion, when we show that the sum F of the series is absolutely continuous (measurability of derivative is established in Lemma 4.3). Take an arbitrary number c > O. By virtue of the uniform convergence of the series FI + F2 + F3 + . .. there exists a number k such that
IIF - (FI
+ ... + Fk)1I
= IIPHI
c
+ FH2 + FH3 + ···11 < "2.
According to the condition, which occurs in the definition of the absolute continuity, find a number 8 > 0 for the function FI + ... + Fk (its absolute continuity follows from Lemma 5.4) and for the number ~ (as c in the condition). For every family, = {(aI, ,6d, ... ,(aj, ,6j)} of pairwise disjoint intervals with db) ~ 8 we have
+ ... + Fk(ai))) - (F(,6i) - (FI (,6i) + ... + Fk(,6i)))I: i = 1, ... ,j} {1(Fk+l (ai) + FH2 (ai) + FH3 (ai) + ... ) - (FHI (,6i) + Fk+2(,6i) + Fk+3(,6i) + ···)1: i =
L {I(F(ai) - (FI (ai) = L
=
L {L {F/(,6i) -
1, ... ,j}
F/(ai): i = 1, ... ,j} : 1 = k + 1, k + 2, k + 3, ... }
~
L {F/(b) -
=
F(b) - (FI(b)
F/(a): 1 = k + 1, k + 2, k + 3, ... } c
+ ... + Fk(b)) < "2.
Therefore
L{IP(ai) - F(,6i)l: i = 1, ... ,j}
+ ... + Fk(ai))) - (F(,6i) - (FI (,6i) + ... + Fk(,6i))) I : {I(FI (ai) + ... + Fk(ai))
~ L{(IP(ai) - (FI(a;)
+L
- (F1 (,6i) ~
c
c
2" + 2" = c.
+ ... + Fk(,6i))l:
i
i = 1, ... ,j}
= 1, ... ,j}
Derivation and integration.
133
This proves the absolute continuity of the function F. The theorem is proved. • Theorem 7.3. Let real functions h, k = 1,2, ... }, be defined almost everywhere on a segment [a, b], a < b, and be Lebesgue integrable. Let their integrals are bounded in totality. Let the sequence Uk: k = 1,2, ... } converge from below to the function f (Le., for almost all x E [a, bJ
f1{X) ~ f2{X) ~ iJ{x) ~ ...
Figure 4.10
and f{x) = limk ..... oo fk{X)) (see Figure 4.10). Then the function f is integrable and b
f
b
f{x)dx =
1~~
a
f
fk{x)dx.
a
Proof. We apply the previous theorem to the series composed by the functions f1' h - fll iJ - f2' f4 - iJ,···· The theorem is proved. • Example 7.3. The characteristic function of an open subset U of a segment [a, b], a < b, may be represented as the sum of the series composed by characteristic functions of its component (see Example 7.2). The integrals of partial sums of this series are bounded from above by the number b - a. Theorem 7.2 implies the integrability of the characteristic function Xu of the set U. Theorems 7.2 and 1.1 imply the equality b
f
Xu{x)dx = J.L{U).
a
So Theorem 7.1 implies the integrability of the characteristic function XM of every closed set M ~ [a, bJ (by virtue of the representation M = [a, bj \ U). By Theorems 7.1 and 1.1 b
f
XM{x)dx
= J.L{M).
a
Example 7.4. Let M be an arbitrary measurable subset of a segment
[a, bj, a < b. The definition of a measurable set implies the existence of a
134
CHAPTER 4
sequence MI ~ M2 ~ M3 ~ ... of subsets of the set M, which are closed in the segment [a, bj and the measure of the union of which is equal to the measure of the set M. The integrals of the characteristic functions XMi are bounded from above by the number b - a. When we apply Theorem 7.3 to the sequences of the characteristic functions, we obtain the integrability of the function XM. By Theorems 7.2, 7.1 and 1.1
!
b
!
boob
XM(x)dx =
a
XM1(X)dx
+ L ! XMk+l\Mk(X)dx
a
=
k=1 a
! b
t; 00
XM1(X)dx+
(
! b
XMk+l(X)dx-
!
b )
XMk(X)dx
00
= J.l(Md
+ L(J.l(Mk+l) -
J.l(Mk ))
k=1 00
k=1
Lemma 7.5. Let real functions f and fo be defined (almost everywhere) and be integrable on a segment [a, b], a < b, of the real line. Let almost everywhere the function fo majorize the function f (i.e., f(x) :::;; fo(x) for almost all x E [a, b]). Then b
! a
b
f(x)dx :::;;
!
fo(x)dx.
a
Proof. Theorem 7.1 and Lemma 7.4 imply:
!
b
a
!
b
fo(x)dx -
a
b
f(x)dx = !Uo(x) - f(x))dx
~ o.
a
The lemma is proved. • Example 7.5. Let a non-negative function f be defined almost everywhere on a segment [a, bj, a < b. Let it have the form mentioned in the Corollary of Theorem 2.3. Let an integrable function fo (defined almost everywhere on the segment [a, b]) majorize almost everywhere the function f. The function f may be represented as the sum of the series alXI + a2X2 + a3X3 + ... , where Xk is the characteristic function of the set Mk and the coefficient ak is non-negative. By Lemma 7.5 the integrals of the partial sums of the mentioned series are bounded in totality by the j
Derivation and integration.
135
integral of the function fo. Theorem 7.2 implies the integrability of the function f. Example 7.6. Let a non-negative measurable function f be defined (almost everywhere) on the segment [a, bj, a < b. Let a Lebesgue integrable real function fo (defined also almost everywhere on the segment [a, b]) majorize almost everywhere the function f. For every pair of positive integers i,j put Mii = {x: x E [a, b], 2- i (j - 1) ~ f(x) < 2- i j}. For i = 1,2, ... define the function fi by the condition
These functions are studied in Example 7.5. Their integrals are bounded from above by the integral of the function fo (see Lemma 7.5). Theorem 7.3 implies the integrability of the function f. Lemma 7.5 implies the inequality b
J
f(x)dx
a
b
~
J
fo(x)dx.
a
Example 7.7. Let a real function f be defined (almost everywhere) on the segment [a, b], a < b, and be Lebesgue integrable. Let f+ be its non-negative part (see §2). When we apply Lemma 6.1 to (every) primitive of the function f by virtue of the nonnegativity of the derivative of a nondecreasing function (everywhere, where the derivative exists) we obtain a representation of the function f as the difference of two non-negative integrable functions: f = fl - f2' Evidently f+(x) ~ fl(X) for almost all x E [a, bj. By remarks of Example 7.6 the function f+ is integrable. Theorem 7.1 implies the integrability of the functions f- (see §2) and If I· If g is one more integrable function, then the functions min{f(x), g(x)} and max{f(x),g(x)} are integrable (see a corresponding representation in §2). The examples mentioned give a rather large reserve of integrable real functions and the remark following the definition of the Lebesgue integral gives a large reserve of integrable vector functions. Lemma 7.6. Let a measurable vector function f with values in the space ]R.n and a non-negative function fo be defined (almost everywhere) on a segment [a, bj. Let IIf(x)1I ~ fo(x) for almost all x E [a, bj. Let the function fo be Lebesgue integrable on the segment [a, bj. Then the function f is also Lebesgue integrable on the segment [a, bj and b
J
f(x)dx
a
~
b
J
fo(x)dx.
a
136
CHAPTER 4
Proof. By our remarks the question about the integrability reduces to the scalar case. By virtue of the reasoning of Example 7.7 in the scalar case the non-negative and nonpositive parts of the function fare integrable. This implies the integrability of the function f. To prove the inequality fix a rectangular system of coordinates in the space ~n in order to make one of its axis co-directed with the vector F(b) - F(a), where F is an arbitrary primitive of the function f (if F(b) - F(a) = 0, then our assertion is obvious). Let fi be the corresponding coordinate function of the function f and ft be its nonnegative part. We have:
J
10
-"-'-- .....
-/0
Figure 4.11
b
IIF(b) - F(a)11 =
J b
fi(X)dx
~
a
ft(x)dx
a
J b
~
fo(x)dx.
a
The lemma is proved. • Theorem 7.4 (Lebesgue). Let functions fk' k = 1,2, ... , f with values in ~n be defined almost everywhere on a segment [a, b], a < b, of the real line. Let the functions fk' k = 1,2, ... , be almost everywhere majorized in the norm by an integrable scalar function fo (i.e., Ilfk(X)11 ~ fo(x) for almost all x E [a, b] and for every k = 1,2, ... ). Let fk(X) ~ f(x) for almost every x E [a, b]. Then the function f is integrable and b
J
f(x)dx =
a
b
l~~
J
h(x)dx.
a
Proof. The integrability of the function f follows from Theorem 2.4 and Lemma 7.6. By virtue of our remarks and Lemma 2.1.2 the question about the limit passage under the integral reduces to the scalar case. In this latter case for k = 1,2, ... consider the functions
137
Derivation and integration. and
cp;;(x) = infUdx), fk+l (x), ... } = lim min{h(x), ... ,fk+p(X)}. p-+oo
Theorem 2.4 and Lemma 7.6 imply the Lebesgue integrability of these functions. Their integrals are majorized in the norm by the integral of the function fo. For almost all x E [a,b] the sequences {cpt(x): k = 1,2, ... } and {cp;; (x): k = 1, 2, ... } converge monotonically to f (x). By Theorem 7.3 b
(7.1)
J b
J
f(x)dx = lim
k-+oo
CPt (x)dx = lim
k-+oo
a
a
J b
cp;;(x)dx.
a
For almost all x E [a, b] we have:
cp;;(x) ~ fk(X) ~ cpt(x). By Lemma 7.5: b
b
J
cp;;(x)dx
~
a
b
J
fk(X)dx
~
a
f
cpt(x)dx.
a
Now (7.1) gives the limit passage in question. The theorem is proved. • Lemma 7.7 (Fatou). Let non-negative functions fk' k = 1,2, ... , be defined almost everywhere on a segment [a, b], a < b, of the real line and be Lebesgue integrable. Let their integrals (over the segment [a, b]) are not greater than m < 00. Let a function f be defined almost everywhere on the segment [a,b] and fk(X) -+ f(x) for almost every x E [a,b]. Then the function f is Lebesgue integrable and b
o~
f
f(x)dx
~ m.
a
Proof. For k = 1,2, ... define the function
CPk(X) = inf{h(x), fk+l(X), ... } = lim minUdx), ... , fk+p(x)}. p-+oo
Theorem 2.6 and Lemma 7.6 imply the Lebesgue integrability of these functions. For almost every x E [a, b] the sequence {CPk(X): k = 1,2, ... } Converges from below to f (x). Since
o~
b
J
CPk{x)dx
a
~
b
J
h(x)dx
a
~ m,
CHAPTER 4
138
for every k = 1,2, ... (see Lemma 7.5), Theorem 7.3 and Lemma 7.5 imply our assertion. The lemma is proved. •
8. Density points and approximate derivatives Let M be an arbitrary measurable subset of a segment [a, b], a < b, X be its characteristic function and H be a primitive of the function x. A point x E [a, b] is called a density point of the set M if the derivative of the function H at the point x exists and is equal to 1. The definition of a primitive and the integrability of the characteristic function of a measurable set (see Example 7.4) imply that almost every point of the set M is its density point and the measure of the set of density points of the set M lying outside M is equal to zero. Fix a point x E [a, b]. For arbitrary real number e denote by /::"0 the segment with endpoints x and x + e. Since the integral of the characteristic function of a (bounded) set is equal to the measure of this set (see Example 7.4), the derivative of the function H at the point x is calculated as the limit of the expressions
as e ~ O. Thus a point x is a density point of the set M if and only if one of the following two equipotent conditions lim JL(/::"o n M) = 1 0-+0
lei
holds. The following fact is essential for us. Lemma 8.1. Let x be a density point of two measurable sets M I , M2 (lying in a segment [a, b], a < b, of the real line). Then the point x a density point of the set MI n M 2 • Proof. For every real number e
Since x is a density point both of the sets MI and M 2 , the terms of the last expression tend to zero as c ~ O. This gives what was required. The lemma is proved. •
Derivation and integration.
139
Lemma 8.2. Let sets M 1 , M2 (lying in a segment [a, b], a < b, 01 the real line) be measurable. Let x E Ml n M2 be a density point lor each 01 them. Let a lunction I : [a, bJ --t IRn have the derivatives at the point x with respect to each 01 these sets. Then the derivatives 01 the lunction I at the point x with respect to the sets Ml and M2 coincide. Proof. By Lemma 8.1 x is a density point of the set M = Ml n M2 too. A density point is always a limit point of the corresponding set. Lemma 4.1 implies the existence of the derivative I~(x) and the equalities I~(x) = I~I (x) and I~(x) = 1~2(X). This gives what was required. The lemma is proved. • A vector v E IRn is called an approximate derivative of a function I : [a, bJ --t IRn at a point x E [a, b], a < b, if there exists a measurable set M ~ [a, bJ such that x EM, the point x is a density point of M and the vector v is equal to the derivative of the function I at the point x with respect to the set M. (When I is a scalar function values of its approximate derivative are numbers.) By Lemma 8.2 the approximate derivative (if it exists) is defined uniquely. In what follows we use the notation of the usual derivative to denote the approximate derivative. We omit the word 'approximate', when this does not lead to misunderstandings. 9. Generalized absolutely continuous functions and the Denjoy integral We say that a function I (defined on a subset M of the real line) is generalized absolutely continuous if it is continuous and the set M may be represented as the union of an at most countable number of subsets, on each of which the function I is absolutely continuous. By Lemma 5.6 we can assume in addition that the subsets in this definition are closed. Lemma 9.1. Let I = 11 U 12 be segments 01 the real line. Let a lunction f with values in the space IRn be defined on the segment I. Let the functions fill and fl12 be generalized absolutely continuous. Then the function f is generalized absolutely continuous too. Proof. The proof is obvious. • Lemma 9.2. Let functions I and 9 with values in the space IRn be defined and be generalized absolutely continuous on the subset M of the real line. Let A be a real number. Then the functions f + g and AI are generalized absolutely continuous. Proof. Let M = u{Ml: i = 1,2, ... } = u{Ml: i = 1,2, ... }. Let the function f be absolutely continuous on every set Ml, i = 1,2, .... Let the function g be absolutely continuous on every set Ml, i = 1,2, .... By
140
CHAPTER 4
Lemma 5.4 the function f+g is absolutely continuous on every set MlnMJ, i,j = 1,2, .... The equality M = u{Ml n MJ: i,j = 1,2, ... } implies the generalized absolute continuity of the function f + g. The generalized absolute continuity of the function )..f follows from Lemma 5.4 directly. The lemma is proved. • By Theorem 6.1 (and Lemma 5.7) a generalized absolutely continuous function has an approximate derivative almost everywhere. By Lemma 4.3 and Theorem 2.3 an approximate derivative is a measurable function. By Lemma 5.3 a vector function with values in the space JRn is generalized absolutely continuous if and only if its coordinate functions are generalized absolutely continuous. By remarks of §4 coordinates of the approximate derivative of such a vector function are approximate derivatives of its corresponding coordinate functions. Lemma 9.3. Let the approximate derivative of a generalized absolutely continuous function f : [a, bj -+ JR, a < b, be non-negative at almost every point of the segment [a, bj. Then the function f is non-decreasing. Proof. Denote by , the set of all connected open subsets of the segment [a, bj, on which the function f is non-decreasing. Let U = U,. If the intersection of two elements of, is nonempty then their union belongs also to ,. So every connected component of U belongs to " Le., the function f is non-decreasing on every of them. The lemma will be proved when we check that [a, bj = U. Assume the opposite, Le., that the set M = [a, bj \ U is nonempty. Let the segment [a, bj be represented as the union of its closed subsets M i , i = 1,2, ... , on every of which the function f is absolutely continuous. We have M = U{M n M i : i = 1,2, ... }. The set M is compact (see Theorem 1.6.2). Hence it is complete (Theorem 1.6.6). Apply Corollary of Baire theorem 1.6.7. By virtue of the definition of the induced topology we obtain the existence of a nonempty interval (a, (3) ~ [a, bj and of a number i = 1,2, ... such that 0 i= (a, /3) n M ~ Mi. At almost every point of the set Mi the approximate derivative coincides with the derivative with respect to this set. When we apply Lemma 7.2 to the function fl[o:,,6] we obtain: (a, /3) E f. This contradicts the nonemptiness of the set (a, /3) n M. Thus our assumption is false and the lemma is proved. • Lemma 9.4. Let the approximate derivative of a generalized absolutely continuous function f : [a, b] -+ JRn, a < b, is equal to zero almost everywhere. Then the function f is constant. Proof. The proof (by virtue of the previous lemma) repeats the proof of Lemma 7.3. • Let a function f be defined almost everywhere either on a segment or on a half interval or on an interval I of the real line and take values in the space JRn. The function f is called Denjoy integrable if there exists a generalized absolutely continuous function F : I -+ JRn such that F'(t) = J(t) for almost i
Derivation and integration.
141
all tEl. The function F (as in the cases of the Riemann and Lebesgue integrals) is called a primitive or an indefinite integral of the function f. As in the cases of the Riemann and Lebesgue integrals (and by analogous arguments, see §7 and Lemma 9.4) a primitive F is defined up to an additive constant and for s, tEl the difference F(t) - F(s) does not depend on the choice of the primitive F. It is called a (definite) integral of the function f from s to t and it is denoted by t
J
f(x)dx.
s
If u is an arbitrary third point of I then t
u
J
f(x)dx
=
u
J
f(x)dx
+
J
f(x)dx.
s s t
In this case too, when we consider the integral as a function of the upper limit of the integration the integral is one of primitives, namely it is the primitive whose value at the lower limit of the integration is zero. Example 9.1. Every Lebesgue integrable function is Denjoy integrable and the notions of primitives and integrals for it in both sense coincide. Example 9.2. Define on the segment [-1, 0] a function f in the following way. Let f(O) = o. For k = 1,2, ... and for an arbitrary point x of the half interval [_2- k +1, _2- k ) put (Figure 4.12) f(x) = (_2)k Ik. The function f is Lebesgue integrable on every segment [-I,e], where -1 < e < O. We can consider its primitive t
F(t) =
J
f(x)dx.
-1
Since IF( _2- k +1) - F( -2- k )1 = 11k --t 0, in view of the alternation of sign of the function on adjacent half interval [_2-k+l, _2- k ) the limit of the function F as t --t 0 exists.
Figure 4.12
142
CHAPTER 4
Extend the function F by the continuity on the entire segment [-I,Oj. Keep its old notation. On every segment [-1, _2- k j, k = 1,2, ... , the function F is a primitive of a Lebesgue integrable function. Thus it is absolutely continuous there. It is also absolutely continuous on the one point set {O}. Thus the function F is generalized absolutely continuous (on the segment [-1,0]). Hence the function f is Denjoy integrable (on the segment [-1,0]) and F is its Denjoy primitive. The function f is not Lebesgue integrable: If we assume such an integrability, we obtain that the function If I must be integrable too. By virtue of the continuity of its primitive the (finite) limit o
lim! If(x)1 dx.
0 ..... 0
-1
must exist. A direct calculation shows that this limit is infinite: 2k
!
If(x)ldx
1
1
= 1 + 2" + ... + k
--+ 00.
-1
Lemma 9.5. The derivative of a non-decreasing continuous real function (defined on a segment [a, b], a < b, of the real line) is Lebesgue integrable. Proof. The hypotheses of Lemma 3.2 hold for the domain M = [a, bj of the function F in question. Lemma 3.2 implies that the derivative f of the function F is defined almost everywhere on the segment [a, bj. Extend the function F on the entire line by putting F(t) = F(a) for t < a and F(t) = F(b) for t > b. Next, put
fork=I,2, ... andtE[a,bj. The definition of a derivative and its existence for the function F almost everywhere on the segment [a, b]) imply that
for almost all t E [a, bJ. The non-decreasing of the function F implies that the functions
143
Derivation and integration. Do it:
i~.(x)dx ~ CT ~ CT ~
2'
F(x)dx -
2'
F(x)dx -
i 7'
F(X)dX)
F(X)dX)
« 00).
F(b) - F(a)
The lemma is proyed. • Remark 9.1. Return to last inequality of the proof of Lemma 9.5. For every k = 1,2, ...
J b
q,k(x)dx
~
F(b) - F(a).
a
When we pass to the limit as k
---t
J
00 we obtain
b
f(x)dx
~
F(b) - F(a).
a
The example of the 'Cantor stairs' shows that the inequality may be strong. Theorem 9.1. A non-negative Denjoy integrable function is Lebesgue integrable. Proof. Let a non-negative function f be defined almost everywhere on the segment la, bj, a < b. Let F be its Denjoy primitive. By Lemma 9.3 the function F is non-decreasing. By Lemma 3.2 at almost every point t of the segment la, bj the function F has a derivative g(t) (in usual sense). By Lemma 8.2 for almost all tEla, bj we have the coincidence f(t) = g(t). By Lemma 9.5 the function g is Lebesgue integrable. Hence the function f is Lebesgue integrable too. The theorem is proved. • The following assertion is analogous to Lemma 7.5. Lemma 9.6. Let real functions f and fo be defined (almost everywhere) and be Denjoy integrable on a segment la, b], a < b, of the real line. Let almost everywhere the function fo majorize the function f. Then b
J
f{t)dt
a
~
b
J
fo(t)dt.
a
•
CHAPTER 4
144
10. Mean Value theorem Lemma 10.1. Let a function f be defined almost everywhere on a segment [a, b] of the real line and take values in the space ~n. Let it be Denjoy integrable (respectively, Lebesgue integrable). Let g : ~n - t ~ be a linear functional. Then the function gf is Denjoy integrable (respectively, Lebesgue integrable) on the segment [a, b] and 9
(!
f(t)dt)
~
i
g(J(t))dt.
Proof. Let F be an arbitrary primitive of the function f. Let f(t) = (f1 (t), ... , fn(t)) and F(t) = (F1 (t), ... Fn(t)). Let g(u) = /-t1X1 + ... + /-tnXn for u = (Xl"'" Xn). We have Ig(u)1 :::;; M sup{lxd, .. ·, Ixnl}, where M = 1/-t11 + ... + l/-tnl. Thus Ig(u)1 :::;; Mllull, i.e., the functional g satisfies the Lipschitz condition. Lemma 5.1 implies the generalized absolute continuity (respectively, absolute continuity) of the function H = gF. For almost all s E [a, b]
H'(s) = (/-t1F1 + ... + /-tnFn)'(s) = /-t1 F; (s) + ... + /-tnF~ (s ) = /-td1(S) + ... + /-tnfn(s) = g(f(s)). Thus b
J
g(f(t))dt = H(b) - H(a) = g(F(b)) - g(F(a))
a
~ g(F(b) -
F(a))
~9
(!
f(t)dt) .
The lemma is proved. • Theorem 10.1. Let a function f be defined on a subset of full measure of a segment [a, b], a < b, of the real line and take values in the space ~n. Let it be Denjoy integrable on the segment [a, b]. Then
b~ a
b
J
f(t)dt E cc(f(M)).
a
Proof. Assume the opposite. By Theorem 2.4.2 there exist a linear functional g : ~n - t ~ and numbers p < q such that (10.1)
g(cc(f(M)))
~
(-oo,pj,
Derivation and integration.
145
9(b~a! !(t)dt) ~q
(10.2) By (10.1) g(f(M))
~
(-oo,p], i.e., g(f(t))
~
p for all t E M. By Lemma
9.6
J b
g(f(t))dt
(10.3)
~ p(b -
a).
a
By (10.2) and Lemma 10.1
b~ a
b
J
~ q,
g(f(t))dt
a
J b
g(f(t))dt
~
q(b - a).
a
This estimate and (10.3) imply the inequality q(b - a) ~ p(b - a). So q ~ p. This contradicts the choice of the numbers p and q. Thus our assumption is false. The theorem is proved. • Notice that in the case of a continuous real function f defined on the entire segment [a, b] the proved inclusion turns up to a little weakened version of the classical Lagrange formula
b~ a
b
J
f(t)dt = f(e) for some e E [a, b].
a
11. Change of variable in an integral and derivation of a composite function Fix notation. Let -00 < a < b < 00, -00 < e < d < 00. Let f: [a, b] --+ [e, d] be a non-decreasing absolutely continuous function. Lemma 11.1. Let a set A ~ [a, b] be measurable and B = f(A). Then the set B is measurable too and p,(B) = fA f'(t) dt. Proof. 1. If A is either an interval (p, q), either a half interval [p, q) or (p, q], or a segment [p, q] then our assertion is obvious: q
p,(B)
= f(q) - f(p) =
J
f'(t) dt.
p
146
CHAPTER 4
II. An arbitrary open set A ~ [a, b] may be represented as the union of the family 'Y of its connected components. By I and Theorems 1.1 and 7.3 (see Example 7.3)
p.(B) = p.(U{J(r): f E 'Y}) ="L{p.(f(f)): fE'Y} =
L ([ f(t) dt:
=
J
r
E'Y}
f'(t) dt.
A
la, b] \ A is open. By II and
III. In the case of a closed set A the set C = Theorem 1.1
J
J
J
a
C
A
b
p.(B)
= (d - c) - p.(f(C)) =
f'(t) dt -
f'(t) dt =
IV. Let us go on to the general case. For arbitrary c that
la, b]
f'(t) dt =
"L {J(sup J) -
> 0 fix 0 > 0 such
and p.(M)
< 0,
f(inf J): J E 'Y}
< c,
if M is an open subset of the segment
J
f'(t) dt.
then
M
where 'Y is the set of all connected component of the set M. This may be done by virtue of the absolute continuity of the function f. Let A be an arbitrary measurable set. By virtue of the measurability of the set A we can find a closed and an open sets Hand G such that
H ~ A~ G The set f(H)
~
and
p.(G \ H) <
o.
B is closed. The set f(G) 2 B is measurable. By II
p.(f(G) \ f(H))
~ p.(f(G \
H)) =
J
f'(t) dt < c.
G\H
By virtue of the measurability of the set f(G) there exists an open set G l ;2 f(G) such that p.(G l \ f(Gd) < ~. Then (11.1)
p.(G l
\
f(H» ~ p.(G l
\
f(G»
+ p.(f(G) \ i
f(H)) <
c
c
2" + 2" = c.
147
Derivation and integration.
The closedness of the set f(H), the openness of the set G I , (11.1) and the arbitrariness in the choice of e imply the measurability of the set B. We have
/-l(J(H))
=
J
J'(t) dt
~
J
J'(t) dt
~
J
J'(t) dt
= /-l(J(G))
H A G
and
/-l(J(H)) ~ /-l(B) ~ /-l(J(G)). So
J
J'(t)dt - /-l(B) <
~ < e.
A
In view of the arbitrariness in the choice of e the last estimate implies the equality
J
J'(t)dt = /-l(B).
A
The lemma is proved. • Lemma 11.2. Let B ~ [c, d]' /-l(B) = 0 and A = f-I(B). Then the measure of the set {t: tEA, the derivative l' (t) exists and l' (t) =1= O} is equal to zero. Proof. The condition /-l(B) = 0 implies the existence of a sequence of open sets G I = [c, dJ ;2 G 2 ;2 G 3 ;2 ... ;2 B such that /-l(G i ) -+ O. Let G = n{Gi : i = 1,2, ... }. For every i = 1,2, ... the set Hi = f-I(G i ) is open in the segment [a, b] and the set (11.2) is the preimage of the set G. By (11.2) and Theorem 1.3 the set H is measurable. By Lemma 11.1
J
J'(t) dt = /-l(G) = O.
H
So 1'(t) = 0 for almost all t E H. Now the inclusion A ~ H implies What was required. The lemma is proved. • Lemma 11.3. Let a function g : [c, d] -+ IRn be Lebesgue integrable. Then there exists a sequence gk : [c, d] -+ IRn, k = 1,2, ... , of continuous functions converging to the function g almost everywhere.
148
CHAPTER 4
IT outside the segment [c, d]. Put
Proof. Extend the function 9 by
Now the derivation of the integral with respect to its upper limit of inte• gration gives what was required. The lemma is proved.
Lemma 11.4. Let a real function 9 : [c, d] --+ lR be measurable and Ig(x) I ~ M for all x E [c, d]. Then the sequence {gk: k = 1,2, ... } in Lemma 11.3 may be chosen in such a manner that Igk(X)1 ~ M for all x E [c, d] and k = 1,2, .... Proof. Let {g,i;: k Put
= 1,2, ... } be sequence pointed in Lemma 11.3. if gic(x) if - M if g,i;(x)
~ ~ ~
M, g,i;(x) -M.
~
M,
We obtain the sequence with the needed properties. The lemma is proved.
•
Theorem 11.1. Let a function 9 : [c, d] --+ lRn be Lebesgue integrable on the segment [c, d] = f([a, b]). Then the function g(f(t» . f'(t) is Lebesgue integrable and d
b
J
J
e
a
g(x) dx =
g(f(t» . f'(t) dt.
Proof. The integration is made in each coordinate independently. Thus the general vector case reduces to the scalar one. In what follows we assume that n = 1. 1. Consider the case of a continuous function g. Let (11.3) Xo = c < Xl < ... < Xp-l < xp = d, mk = infg([xk_I,Xk]), Mk = sUpg([Xk_I,Xk]) for k = 1, .. . p, tk = j-I(Xk) for k = 0,1, ... p,
149
Derivation and integration. By Theorem 7.5
L{mk(Xk - xk-d: k = 1, .. . p}
7
= L{mk
k= 1, ... P}
f'(t) dt:
tk-l
J b
~
(11.4)
g(f(t)) . f'(t) dt
a
~L
7
{Mk
f'(t) dt: k = 1, .. .
p}
tk-l
Let 8 = SUp{ Xl - XO, ... ,Xp - Xp-l}. By Theorem 7.4 the end sides in (11.4) tend to fed g(x) dx as 8 --t 0 in (11.3). This gives what was required. II. Consider the case of a measurable function g. Assume that the function 9 is bounded in modulus by a number M: Ig(x)1 ~ M for all X E [c, d]. Fix an arbitrary sequence of continuous functions {gk: k = 1,2, ... } in a manner that the absolute values of the functions are bounded by the number M and the sequence converges almost everywhere to the function g, see Lemma 11.4. By I d
J
g(x) dx =
e
d
kl~~
b
J
gdx) dx =
kl~~
J
gk(f(t)) . f'(t) dt.
a
c
The expressions in the integral in the last side is almost everywhere bounded from above by the Lebesgue integrable function M f'(t). It remains to show that almost everywhere
gk(f(t)) . f'(t)
--t
g(f(t)) . f'(t).
The derivative f'(t) exists on the set M ~ [a, bj of the measure b - a. The condition gk(X) --t g(x) may be not fulfilled on the set N ~ [c, dj of measure zero. At points of the set M n j-l([C, dj \ N) = M \ j-l(N) we have the needed convergence. We will get what was required when we show that the convergence (11.5) takes place for almost all t E M n j-l(N). By Lemma 11.2 f'(t) = 0 for almost all t E j-l(N). This implies immediately what was required.
CHAPTER 4
150
III. Consider the case of a non-negative integrable function g. Omit the assumption of II about its boundedness. Let gk(X) = min{g(x), k}. We have gk(X) - t g(x) for every x E [c, dJ. Hence the convergence (11.5) takes place at every point t E [a, b], at which the derivative f'(t) exists. By II d
d
! 9(X) dx
b
= k-+oo lim !9k(X)
e
dx
= k-+oo lim !9k(J(t». J'(t)
dt.
a
e
Values of the integrals in the last side constitute a monotonically increasing g(x) dx. sequence bounded from above by the number Theorem 1.7.3 implies that the function g(J(t» . f'(t) is integrable and
t
!
d
!
b
g(x) dx
=
g(J(t» . J'(t) dt.
a
e
IV. An arbitrary function 9 may be represented as the difference of two non-negative integrable functions: 9 = gl - g2. When we apply III to each of the functions gl and g2, we obtain d
! e
d
g(x) dx =
!
d
91(X) dx -
!
e
g2(X) dx
e b
b
= j gl(J(t))· J'(t) dt - j g2(J(t» . J'(t) dt a
a b
= j(91(J(t» - g2(J(t»)· J'(t) dt a b
= j g(J(t» . J'(t) dt. a
The theorem is proved. • Corollary. Let under the hypotheses of Theorem 11.1 G be a primitive of the function g. Then for almost all t E [a, bJ (11.6)
(G(J(t»)' = g(J(t)) . J'(t).
This may be obtained by the derivation of the expression t
G(f(t» - G(f(a» =
Jg(f(s» . f'(s) ds a
151
Derivation and integration.
with respect to t. • Remark 11.1. Let a function G be absolutely continuous on the closed subset M of the segment [c,d] and Ml = f-l(M}. Then the function Gf is absolutely continuous on the set MI' The proof of this fact is quite obvious. The last observation implies immediately that if the function G is generalized absolutely continuous, then the function G f is also generalized absolutely continuous. Assertion 11.1. Let a function g be Denjoy integrable on the segment [c, d] = f([a, b]) and G be a primitive of the function g. Then for almost all t E [a, b] condition (11.6) holds. Proof. The function G is generalized absolutely continuous. Thus the segment [c, d] may be represented as the union of countable number of closed subsets M i , i = 1,2, ... , on each of which the function G is absolutely continuous. Fix i = 1,2, .... Let 'Y be a family of all connected components of the complement [c, d] \ Mi' The function G(t) {
H(t) =
t-
in~f
G(supf)
supf - mff (Figure 4.13)
+
supf -: t G(infr) supf - mff
for t E f E 'Y
is absolutely continuous. It is a primitive of the function
'-v-'
r
Figure 4.13
Figure 4.14
get) h(t) = {
G(supr) - G(inff) supf - inff
for
t E f E 'Y.
(Figure 4.14). By Remark 11.1 at almost every point t of the set D = f-l(Mi ) the derivative (H(f(t)))'v with respect to set D exists. It is equal to the derivative (G(f(t)))'v. Almost all points of the set Dare
CHAPTER 4
152
density points of this set. At these points the equality goes into the equality of the corresponding approximate derivatives. By Corollary of Theorem 11.1 and by Lemma 11.2
(G(f(t)))'
= (H(f(t)))' = h(f(t)) . f'(t) = g(J(t)) . f'(t).
Since this is true for every i = 1,2, ... the assertion is proved.
•
12. Measurable multi-valued mappings Let a multi-valued mapping F be defined on a measurable subset M of the real line and take values in a topological space Y. The mapping F is called measurable if for every open subset U of the space Y the set F-l(U) = {t: t E M, F(t) n U 1= 0} is measurable. In the case of a single valued mapping this notion of the measurability coincides with this one of
§2. Theorem 12.1. Let a multi-valued mapping F be defined on a measurable subset M of the real line and take compact values in a metric space Y. Then the following conditions are equipotent: a) the mapping F is measurable; b) for every closed subset H of the space Y the set F- 1 (H) is measurable; c) for every open subset U of the space Y the set {t: t E M, F(t) ~ U} is measurable; d) for every closed subset H of the space Y the set {t: t E M, F(t) ~ H} is measurable. If Y is the Euclidean space then the listed conditions are equipotent to the conditions: e) for every open and convex subset U of the space Y the set F- 1 (U) is measurable; f) for every open ball U of the space Y the set F-l(U) is measurable. Proof. a:::}b. By virtue of the compactness of values of the mapping F and Lemma 2.3.8 we have F-1(H) = n'f:lF-1(Ol/kH). Now a) and Theorem 1.3 imply the measurability of the set F-l(H). b:::} a. Let U be an open subset of the space Y. For k = 1,2,... put Hk = Y \ Ol/k(Y \ U). The set H k, k = 1,2, ... , is closed, U = U'f:lHk and F-l(U) = U{F-l(Hd: k = 1,2, ... }. Now b) and Theorem 1.1.2 implies the measurability of the set F-l(U). a ¢:>c and b ¢:>d. It is obvious because
{t: tEM, F(t)c;H}=M\F-1(Y\H). a:::}e:::}j. It is obvious. j:::}a. Open balls constitute a base of the space ~n. Therefore every open subset U of the space ~n may be represented as the union of a family
Derivation and integration.
153
'Y of open balls. By Theorem 1.6.10 the presence of a countable base of the space ~n allows to assume that the family 'Y is (at most) countable: 'Y = {Vk : k = 1,2, ... }. The equality
and Theorem 1.1.2 imply the measurability of the set F-I(U). The theorem is proved. • Theorem 12.2. Let a set M ~ ~ be represented as the union of an {at most} countable family of measurable sets {Mk: k = 1,2, ... }. Let a multi-valued mapping F be defined on the set M, take values in a topological space Y and be measurable on every set Mk {i. e., the mapping F IMk , k = 1,2, ... , be measurable}. Then the mapping F is measurable. Proof. The proof repeats the proof of Theorem 2.3 (the assertion in question is a direct generalization of the theorem). • Lemma 12.1. Let Fk : M - t Y, k = 1,2, ... , be measurable multivalued mappings of a measurable subset M of the real line in a metric space Y. Let their values be compact, FI(t) ;2 F 2 (t) ;2 F3(t) ;2 ... and F(t) = n{Fk(t): k = 1,2, ... } for every point t E M. Then the mapping F is measurable. Proof. Let H be an arbitrary closed subset of the space Y. By Corollary of Theorem 1.6.3 the compactness of values of the mappings under consideration implies the equality F-I(H) = n{Fk-I{H): k = 1,2, ... }. Theorems 1.3 and 1.12b imply the measurability of the set F- I (H). By Theorem 1.12b the mapping F is measurable. The lemma is proved. • Lemma 12.2. Let a multi-valued mapping F be defined on a measurable subset M of the real line, be measurable and take values in a metric space Y. Let HI be a closed subset of the space Y and FI (t) = F{t) n HI for t EM. Then the mapping FI : M - t Y is measurable. Proof. For every closed subset H of the space Y the set F1-l{H) = F-I{H n Hd is measurable by Theorem 12.1b. One more application of this theorem gives the measurability of the mapping F I . The lemma is proved. • Theorem 12.3. Let M be a measurable subset of the real line. Let Y be a metric space with a countable base. Let values of a (multi-valued) mapping F : M - t Y be nonempty and compact. Then there exists a single valued measurable mapping f : M - t Y such that f{t) E F{t) for every point
tEM. Proof. Construct a sequence {Fk: k = 0, 1,2, ... } of multi-valued measurable mappings. Put Fo = F. Let the mappings Fk - 1 , k = 1,2, ... , be constructed. Construct now the mapping Fk in the following way. By Theorem 1.6.10 the open cover {O(y, 21k): y E Y} of the space Y has an
CHAPTER 4
154
(at most) countable subcover {O(Yj, 21k): j = 1,2, ... }. Let
Ml = Fk-!l M2 = Mj =
([O(Yl' 21k)]) ,
21k)]) \ M1, ... , 1k)]) \ (Ml U··· U Mj-d,··· Fk-!l ([O(Yj, 2
F;;!l
([0(Y2'
.
The measurability of these sets follows from Theorem 12.1b and results of §1. For t E M j put Fk(t) = Fk-1(t) n [O(Yj, A)]. Since M = U{Mj : j = 1,2, ... } and the sets M j , j = 1,2, ... , are pairwise disjoint, the mapping Fk : M - t Y is defined correctly. For every t E M the set Fk(t) is nonempty and compact. If points x, Y belong to Fk(t), then p(x, y) ~ i.e., diamFk(t) ~ By Lemma 12.2 and Theorem 12.2 the mapping Fk is measurable. For t E M put F(t) = n{Fk(t): k = 1,2, ... }. By Lemma 12.1 the mapping F is measurable. By virtue of the compactness of the sets Fk(t), k = 1,2, ... , and Corollary of Theorem 1.6.3 the set F(t) is nonempty. Since
h
i.
diamF(t)
~ diamFk(t) ~ ~
for every k = 1,2, ... the set F(t) is one point. Denote it by f(t). The mapping f is constructed. The theorem is proved. • We will need the following additional assertions. Lemma 12.3. Let a multi-valued mapping F be defined on a measurable subset M of the real line, take values in the Euclidean space ~n and be measurable. Let a non-negative measurable function cp be defined on the set M. For t E M let G(t) = Of(F(t),cp(t)). Then the mapping G is measurable. Proof. The proof use Theorem 12.1f. Take an arbitrary point x E ~n and an arbitrary number c > O. The set
n OEX =J 0} F(t) n OE+CP(t)X =J 0}
G-1(OEX) = {t: t E M, G(t) = {t: t E M,
= U{ {t: t E M, F(t)
n Orx =J 0}
n{t: tEM, cp(t»r-c}: rEQ, r>O} is measurable by virtue of the measurability of the mapping F, of the measurability of the function cp and by Theorems 2.8, 1.2 and 1.3. Now our • assertion follows from Theorem 12.1. The lemma is proved. Lemma 12.4. Let measurable multi-valued mappings F, G: M - t JRn be defined on a measurable subset M of the real line JR. Let their values be
155
Derivation and integration.
closed. Let H{t) = F(t) n G(t) for t E M. Then the mapping H : M --+ ]Rn is measurable. Proof. 1. Let U be an arbitrary open subset of the space ]Rn. By virtue of the measurability of the mappings F and G and Theorem 1.3 the set {t: t
M, F(t) n U 1= 0, G(t) n U 1= 0} ={t: tEM, F(t)nU1=0}n{t: tEM, G(t)nU1=0} E
is measurable. II. Let 13 be an arbitrary (at most) countable family of open subsets of the space ]Rn. By I and Theorem 1.2 the set
u{ {t: t
E
M, F(t) n U 1= 0, G(t) n U 1= 0}: U E f3}
is measurable. III. Let {13k: k = 1,2, ... } be an (at most) countable family of (at most) countable families of open subsets of the space ]Rn. By II and Theorem 1.3 the set
n{u{{ t: t
E
M, F(t) n U 1=
0,
G(t) n U 1=
0}:
U
E
f3d: k
= 1,2, ... }
is measurable. IV. Let K be an arbitrary compact subset of ]Rn. Show that the set
M*={t: tEM, H(t)nK1=0} is measurable. Let 13 be an arbitrary countable base of the space ]Rn. For k = 1, 2, ... let 13k = {U: U E 13, Un K 1= 0, diamU < 2- k}. Evidently M* ~ M**, where M** = n{u{ {t: t E M, F(t) n U 1= 0, G(t) n U 1= 0}: U E f3d: k = 1,2, ... }. Lemma 2.3.8 implies the inverse inclusion M** ~ M*. Thus M* = M**. By III the set M* is measurable. V. Let x be an arbitrary point of the space ]Rn and c > O. The representation O(x, c) = U {Of (x, c(k;
1)) : k
= 2,3, ... }
implies the equality H-l(O(X, c»
= {t: t
E
= U {{ t:
M, H(t) n O(x, c) 1= t E M,
H(t) n Of (x,
0}
c(k-l») k 1= 0 } :
k
= 1,2, ..
CHAPTER 4
156
By virtue of the compactness of the sets Of (x,c(k -l)/k), k = 1,2, ... , IV and Theorem 1.2 the set H- 1 (0(x, c)) is measurable. Referring to Theorem 12.1 completes the proof. The lemma is proved. •
13. Multi-valued mappings defined on products Consider the following situation:
(13.1) I = [a, bj, a < b, is a segment of the real line; K is a metric compact; Y is a metric space with a countable base; F : I x K -+ Y is a multi-valued mapping with nonempty compact values; for every tEl the mapping F(t, x) is upper semicontinuous in x; for every x E K the mapping F( t, x) is measurable in t.
With the notation of (13.1) for A
~
K and B
~
Y let us put
M*(A,B)={t: tEl, F({t}xA)~B}.
We will also need the following strengthening of the last condition in (13.1): (13.2) for every open subset U of the compactum K and for every open subset V of the space Y the set M*(U, V) is measurable.
Lemma 13.1. With the notation of (13.1) for every closed subset H of the compactum K and for every open subset V of the space Y we have M*(H, V) = U'f:1M*(OtH, V). Proof. For every k = 1,2, ... we have M*(OtH, V) ~ M*{H, V) because H ~ 0tH. Thus U'f:1M*(OtH, V) ~ M*(H, V). The problem reduces to the proof of the inclusion M*(H, V) ~ Uk'=l M*(OtH, V). Assume the opposite. Let t E M*(H, V) \ U'f:1M*(OtH, V). Since t ~ U'f:1M*(OtH, V) for every k = 1,2, ... there are a point Xk E 0tH and a point Yk E F(t, xd \ By virtue of the compactness of the space K the sequence {Xk: k = 1,2, ... } has a subsequence {Xk: k E A} converging to a point x* E K. Since Xk E 0tH for k = 1,2, ... , x* E H (see Corollary of Lemma 2.3.8). The existence of the sequences {Xk : k E A} and {Yk: k E A}, Theorem 2.5.1 and the inclusion F(t,x) ~ F({t} x H) ~ V imply that the mapping F(t, x) is not upper semicontinuous in x at the point (t, x*) (because the sequence {Yk: k E A} ~ K \ V has no a subsequence converging to a point of F(t,x». This contradicts (13.1). Thus our assumption is false. The lemma is proved. •
v.
Derivation and integration.
157
Lemma 13.2. Under the hypotheses of (13.1) condition (13.2) is equipotent to the condition: (13.3) for every closed subset H of the compactum K and for every open subset V of the space Y the set M*(H, V) is measurable.
Proof. Necessity follows from Lemma 13.1 and Theorem 1.2. Sufficiency. Let U be an open subset of the compactum K and U For k = 1,2, ... put Hk = K \ 0t(K \ U). Since U = Uk:lHk,
=1=
K.
The measurability of the set M*(U, V) follows from (13.3) and Theorem 1.3. Lemma is proved. • Theorem 13.1. Let conditions (13.1) and (13.2) hold. Let X be a closed subset of the product I x K. Let G(t) = F(( {t} x K) n X) for t E I. Then the mapping G : I --t Y is measurable . Proof. I. Let the set X be the product of compacta II and K l . In this case our assertion follows from Lemma 13.2 and Theorem 12.1c. II. Let the set X be the union of the products I j x K j , j = 1, ... , s, of compacta I j and K j . Let a mapping G j be constructed according to the statement of the theorem by the set I j x K j • By I the mapping G j is measurable. Therefore for every open subset V of the space Y the set Gjl (V) is measurable. The equality G- l (V) = Uj=l Gjl (V) implies the measurability of the set G-l(V). Hence the mapping G is measurable. III. Let us go on to the general case. Let k = 1,2, .... For every point x E X find its Tychonoffneighborhood Tx x Ux such that [Tx x Uxl ~ 0tX. The constructed open cover {Tx x Ux : x E X} of the compactum X contains a finite subcover {TXl x UXl , • • • ,Tx. x Ux.}. Put
The set X k satisfies the hypotheses of II. The corresponding mapping G k is measurable. By Theorem 1.7.8 values of the mapping G k are compact. Since X ~ n~lXk ~ n~lOtX = X (i.e., X = nk:lxd, the upper semicontinuity of the mapping F(t, x) in x implies: G(t) = nk:lGk(t). By Lemma 12.1 the mapping G is measurable. • The theorem is proved. Let us highlight a particular case of the theorem proved. Corollary. Let conditions (13.1) and (13.2) hold. Let f be a continuous (single valued) mapping of the segment I in the compactum K.
158
CHAPTER 4
Let the (multi-valued) mapping G : I - t Y be defined by the formula G(t) = F(t, f(t». Then the mapping G is measurable. • Theorem 13.2. Let condition (13.1) hold. Then condition (13.2) is equipotent to the condition:
(13.4) for every number € > 0 there exists a closed subset H of the segment [a, b] such that f..L(H) ~ b-a-€ and the mapping FIHxK is upper semicontinuous.
a Figure 4.15
Proof. Necessity. Fix a countable base a = {Am: m = 1,2, ... } of the compactum K and a countable base f3* of the space Y. The family f3 of the unions of all possible finite subfamily of the base f3* is countable: f3 = {Bn: n = 1,2, ... }. By (13.2) the set Mmn = M*(Am' B n ), m, n = 1,2, ... , is measurable. Therefore there are closed Pmn and open Qmn subsets of the segment [a, b] such that Pmn ~ Mmn ~ Qmn and f..L(Wmn ) < 2- m- n€, where W mn = Qmn \ Pmn- The set [a, b] \ W mn may be represented as the union of two disjoint closed subsets Pmn and [a, b] \ Qmn of the segment [a, b]. Both these sets are closed in the subspace [a, b] \ Wmn- Since they are complements in [a, b] \ W mn of each other they are open in [a, b] \ W mn. By Theorem 1.2 the measure of the set W = U{Wmn: m, n = 1,2, ... } is less than €. Put H = [a,b] \ W. Take an arbitrary open subset V of the space Y. Our aim is to show that the set G v = {(t,x): (t,x) E H x K, F(t,x) ~ V} is open in the product H x K (Figure 4.15). Consider first the particular case V = Bn- If (t* , x*) E G Bn then by virtue of the upper semicontinuity of the mapping F(t, x) in x there exists an element Am 3 x* of the base a such that F( {t*} x Am) ~ B n , i.e., t* E Mmn- In view of the arbitrariness of the point (t*,x*) E G Bn this means that G Bn = U{(Mmn \ W) x Am: m = 1,2, ... }. Hence the set G Bn is open in the product H x K. In the general case for every point (t*, y*) E G v in view of the compactness of the set F(t*, x*) there are elements Vi, ... , Vs of the base f3* such that F( t*, x*) ~ Vi U ... U Vs ~ V. The set Vi U ... U Vs ~ V belongs to the family f3, i.e., G v = U{ G B : B E {3, B ~ V}. This implies the openness of the set G v in the product HxK. Sufficiency. By Lemma 13.2 it is sufficient to show that (13.4) implies (13.3). Let Kl be a closed subset of the compactum K, V be an open {
Derivation and integration.
159
subset of the space Y. For every j = 1,2, ... fix a closed subset H j of the segment [a, b] such that {t(Hj) ~ b-a- j and the mapping FIHj xK is upper semicontinuous. The last condition implies the openness of the set G = {(t, y): (t, y) E H j xK, F(t, y) ~ V} in the product H j x K. Theorem 1.6.2 implies the compactness of the set (Hj x K) \ G. Lemma 2.1.1 and Theorem 1.7.8 and 1.6.1 imply the closed neSS of the set P = {t: t E H j , ({t} x Kd\G =1= 0}. By Theorem 1.4.2 the set H j \P = M(Kl' V)nHj is open in the space H j . Lemma 1.7 implies its measurability. By Theorem 1.2 the set Q = M(Kl' V)n(U{Hj: j = 1,2, ... }) is measurable. We have M(Kl' V)\ Q ~ nk:l ([a, b] \ Hj ). By Lemma 1.1 and Theorem 1.1 the measure of the last set is equal to zero. Theorem 1.1 implies the measurability of the set M(K1 , V). The theorem is proved. • Lemma 13.3. Let condition (13.1) hold. Let:
(13.5) for every point t of the segment [a, b] and for every open subset V of the space Y the set N = {x: x E K, F(t,x)nV =1= 0} lie in the closure of its interior: N ~ [(N)]. Then condition (13.2) holds too. Proof. I. Let U be an open subset of the compactum K, V be an open subset of the space Y. Fix a countable everywhere dense subset A of U ~ K. Evidently the set M = n{M*({x}, V): x E A} contains the set M*(U, V). By Theorem 1.3 the set M is measurable. Show that M ~ M*(U, [V]). Assume the opposite, i.e., the existence of a point t* E M \ M*(U, [V]). Then there exists a point x* E U such that F(t*, x*) \ [V] =1= 0. By virtue of the density of the set A in U and of (13.5) F(t*,x) \ [V] =1= 0 for a point x E A. This contradicts the choice of the point t* EM. Hence our assumption is false and the inclusion in question is true. Thus M*(U, V) ~ M ~ M*(U, [V]). II. Let H be a closed subset of the compactum K and V be an open SUbset of the space Y. By Lemma 13.1 M*(H, V) = nk:lM*(OtH, V). By I for every k = 1,2, ... there exists a measurable set Mk such that M*(OtH, V) ~ Mk ~ M*(OtH, [V]) (~ M*(H, [V])). By Theorem 1.2 the set Mo = Uk:lMk is measurable. By Lemma 13.1 M.*(H, V) ~ Mo ~ M*(H, [V]). III. Let H be a closed subset of the compactum K and V be an open SUbset of the space Y.
CHAPTER 4
160
For k = 1,2, ... put Vk = Y\ [Ot(Y\ V)]. The sets Vk , k are open and Vi ~ [Vd ~ V2 ~ ... ~ V (therefore M*(H, Vd
~
M*(H, [Vi])
~
M*(H, V2 )
~
•••
~
= 1,2, ... ,
M*(H, V)),
Uk:l Vk = V (see, for instance, Corollary 2 of Lemma 2.3.1 and Lemma 2.3.2). By Theorem 1.7.8 for every t E [a, b] the set F( {t} x H) is compact. Therefore M*(H, V) = Uk:1M*(H, Vk ).
By II for every k = 1,2, ... we can fix a measurable set Mk such that M*(H, Vk) ~ Mk ~ M*(H, [Vk]). Hence M*(H, V) = Uk:lMk' By Theorem 1.2 the set M*(H, V) is measurable. We have checked the fulfilment of the condition (13.3). Now our assertion follows from Lemma 13.2. The lemma is proved. • Theorem 13.3. Let condition (13.1) hold. Let the mapping F be single valued. Then condition (13.2) holds. Proof. Notice that in the case of a single valued mapping the condition of the upper semicontinuity of the mapping F in the second argument goes into the condition of the continuity and the fulfilment of (13.5) become obvious. By virtue of this observation our assertion follows from Lemma • 13.3. The theorem is proved. Theorems 13.3 and 13.2 imply: Theorem 13.4 (Scorza-Dragoni). Let:
(13.6) I = [a, b], a < b, be a segment of the real line; K be a metric compact; Y be a metric space with a countable base; f : I x K -+ Y be a (single valued) mapping; for every tEl the mapping f (t, x) be continuous in x; for every x E K the mapping f (t, x) be measurable in t.
Then:
(13.7) for every number £ > 0 there exists a closed subset H of the segment [a, b] such that J.t(H) ~ b - a - £ and the mapping fl HxK is continuous . • Theorem 13.5 (A strengthening of the Egorov theorem). Let condition (13.6) hold. Let every function Ii, i = 1,2, ... , satisfy the condition imposed on the function f in (13.6). For almost every t E [a, b] let the sequence of functions gH x) = fi (t, x) converge uniformly to the function lex) = f(t, x). Then for every £ > 0 there exists a closed set H ~ [a, b] such that J.L(H) ~ b - a - £, the functions fil HxK ' i = 1,2, ... , and fl HxK are continuous and the sequence {fiIHxK: i = 1,2, ... } converges uniformly to the function fl HxK '
Derivation and integration.
161
Proof. I. We can change values of the functions fi' i = 1,2, ... , f on a set A x K, where the measure of the set A ~ [a, bj is equal to zero, in order to guarantee the fulfilment of the condition: for every t E [a, b] the sequence of functions {gf : i uniformly to the function l.
= 1,2, ... } converges
II. Denote by Ko the compactum K x ({O} U {2- i Define the function: F(t
x) = {fi(t, y, x) ,y, f(t,y)
if if
x = 2- i i x = 0
i = 1,2, ... }).
= 1,2, ... ,
on the product [a, bj x K o. The hypotheses of Theorem 13.4 hold. This implies that we can fix a closed set Hl ~ [a, b] such that p,(Hd ~ b - a - ~ and the function FIHl xKo is continuous. There exists an open set U 2 A of the measure < ~. The measure of the set H = Hl \ U is greater than b - a-c. Values of the functions fi' i = 1,2, ... , f were not changed on the set H x K. The function FI HxKo is continuous. By Theorems 3.2.4 and 3.2.3 the last fact implies the uniform convergence fi IH x K -+ f IH x K· The theorem is proved. • As a corollary of Theorem 13.5 we obtain: Theorem 13.6 (Egorov). Let real functions h i = 0,1,2, ... , be defined and be measurable on a segment [a, b], -00 < a < b < 00, of the real line IR. Let fi(t) -+ fo(t) for almost all t E [a, b]. Then for every c > 0 there exists a closed subset H of the segment [a, b] such that p,(H) ~ b - a - c, the functions filH' i = 0,1,2, ... , are continuous and the sequence {fiI H : i = 1,2, ... } converges uniformly to the function foln- • When fi == f Theorem 13.6 implies: Theorem 13.7 (Luzin). Let a real function f be defined and be measurable on a segment [a, b]' a < b, of the real line. Then for every number c > 0 there exists a closed subset H of the segment [a, b] such that p,{H) ~ b-a-c and the function flH is continuous. • Remark 13.1. Let us strengthen Lemma 11.3 from the point of view of results of this section. Assume that the conditions (13.1-2) hold for Y = IRn. Following Theorem 13.2 fix a sequence {a, b} ~ Hl ~ H2 ~ ... of closed subsets of the segment [a, b] such that P,{Hi) -+ b - a as i -+ 00 ~nd the mapping F IHi X K is upper semi continuous (see also Theorem 1. 7.6 In connection with the construction of a non-decreasing sequence of sets). Define a mapping Fi : [a, b] x K -+ IRn. Let it coincide with F on the set
162
CHAPTER 4
Hi X K: Fi IHi X K = F IHi X K' Let (c, d) be a connected component of the set [a, b] \ Hi' For t E (c, d) and x E K put t-c Fi(t,x) = d _ cF(d,x)
d-t
+ d _ cF(c,x).
The upper semicontinuity of the mapping Fi may be easily established with the using of Theorem 2.5.1. The convergence of the sequence of the mappings {Fi: i = 1,2, ... } to the mapping F has a rather strong character: for almost all t E [a,b] we have Fi(t,x) = f(t,x) for all x E K, beginning with some i = io. If the mapping F is single valued then the mappings Fi are single valued too. If values of the mapping F are convex, then values of the mappings Fi are convex too. If values of the mapping F are bounded from above by a number M, that is lIuli ~ M for every vector u E F([a, b] x K), then the analogous condition is true for the mappings Fi .
CHAPTER 5
WEAK TOPOLOGY ON THE SPACE L1 AND DERIVATION OF CONVERGENT SEQUENCES
In this chapter we collect the material in Functional Analysis and in the theory of Functions of Real Variable which is important for our Cauchy problem theory for differential equations and inclusions. In particular, we obtain the possibility of investigating of equations with complicated discontinuities in their right hand sides in space variables. Assertions of the functional analysis cited in this chapter will be helpful in our methods of proving the convergence of sequences of solution spaces of equations. In particular, they give the possibility of studying the mentioned class of equations. But they may be applied with advantage in the investigation of equations with continuous right hand sides too. We will meet such situations later. We do not aim here to give an exhaustive account of the stated fields of mathematics, but in the framework of our needs the corresponding material is given in entirety. The Hahn-Banach theorem and a simple version of the Dunford-Pettis criterion of the weak compactness are key points in this chapters (see §8 here and [Ed]). 1. Continuous linear functionals
We will restrict ourselves to the consideration of vector spaces over the field of real numbers JR. Let L be a normed space (see §2.4). Let D = {u: u E L, Ilull ~ I} be the closed unit ball of the space L. Assertion 1.1. Let f : L --t JR be a linear functional. Then the following conditions are equipotent: a) the functional f is continuous; b) the functional f is continuous at the point 0; c) sup{lf(u)l: u E D} < 00. Proof. a=>b. It is obvious. b=>a. Take arbitrary u ELand c > O. By b) there exists a number 6 > 0 such that. if IIwll < 6, then If(w)1 < c. If now v E 06U, then the
164
CHAPTER 5
vector w = v - u belongs to 0 6 0. So
If(v) - f(u)1
= If(v -
u)1
= If(w)1 < c.
b=;.c. By b) there exists a number c > 0 such that if If(w)1 < 1. If now u E D, then
Therefore
c
2"lf(u)1 < 1,
Ilwll <
c, then
2 If(u)1 < -.
c By virtue of the arbitrariness of u E D this implies that
sup{lf(u)l: u E D}
~
2 -. c
c=;.b. Let m = sup{lf(u)l: u ED}. For arbitrary c > 0 consider
8 = 2~. If now u E 0 6 0, then
2m
-u c
E
D,
2m
-If(u)1 c
~ m,
If(u)1
~
c -2 < c.
The assertion is proved. We noticed in §2.4 that every linear functional on the space
• ~n
is con-
tinuous. The set L * of all continuous linear functionals on the space L is a vector space. Assertion 1.1 allows to introduce on L* the function II . II : L* ~ [0, (0), lIull = sup{lf(u)1 :u ED}. We use for the function the symbol of the norm. Show that it is in fact a norm. If f E L* and flD == 0, then f == o. This implies that Ilfll = 0 ifand only if f == o. Thus the new function II . II satisfies condition 1) of the definition of a norm in §2.4. The fulfilment of condition 2) is obvious. Show the fulfilment of condition 3). Let f, gEL and u E D. We have
If(u)
+ g(u)1
~ If(u)1
+ Ig(u)1
~
Ilfll + IIgll·
Since it holds for every vector u E D,
Ilf + gil
= sup{lf(u)
+ g(u)l:
u
E
IIfll + IIgll· If(u)1 ~ Ilfll . Ilull·
D} ~
Assertion 1.2. Let u ELand f E L*. Then Proof. For u = 0 the estimate in question is obvious. Let u # O. The vector v = II~II belongs to D. Therefore
If(v)1
~
Ilfll,
If(u)1
-W~lIfli.
Weak topology on the space L1 and derivation of convergent sequences. 165 This gives what was required. • Besides the topology generated by the mentioned norm the space L * carries a so called 'weak' topology. The 'weak' topology is generated by the 'embedding' i : L* ~ II~.L: each linear functional f E L* is a function f : L ~ JR and so it represents an element of the Tychonoff product JRL. For the normed space L * we will keep the notation L *. Denote the space L * with the weak topology by L:. Associate to an arbitrary vector u E L the linear functional CPu : L * ~ JR, CPu(J) = f(u). It is continuous: (1.1)
if f E L* and
Ilfll
~ 1, then
ICPu(J)1
=
If(u)1 ~
IIfll·llull
~
Ilull·
Remark 1.1. For u E L the mapping CPu : L* ~ JR coincides with the composition of the mapping i and of the projection of the product JRL on the corresponding factor. This implies the continuity of the mapping i (see Theorem 2.1.1). The mapping : L ~ (L*)*, (u) = CPu, is linear:
CPo:u+(3v(J)
= f(au + /3v) = af(u) + /3f(v) = acpu(J) + /3CPv(J).
It is continuous: by virtue of the linearity of and (1.1)
1I(u) - (v) II
=
11(u - v)11
~
Ilu - vii.
In the next section we will continue the investigation of the mapping . Remark 1.2. Often the mapping is an isomorphism. In this case space L is called reflexive. But this need not take place. Assertion 1.3. The space L* is complete. (Independently of the completeness of the space L or its absence.) Proof. I. For every fixed point u E D the mapping BC(D, JR) ~ JR (see notation in §3.1), which associates to a function f E BC(D, JR) its value f(u), is continuous because If(u) - g(u)1 ~ Ilf - gil. II. Because of I for every a E [0, 1J and u, v E D the set
{J: f E BC(D, JR), f(au
+ (1 -
a)v) = af(u)
+ (1 -
a)f(v)}
is closed in the space BC(D, JR). III. The closedness of the set
A
= {J: f
f(au + (1 - a)v) = af(u) for every a E [O,lJ and u, v E D}
E BC(D, JR),
+ (1 -
a)f(v)
in the space BC(D, JR) follows from II and from Theorem 1.3.4. IV. The mapping L* ~ A, which associates to a functional f E L* its restriction fl D , is an isometry.
166
CHAPTER 5
V. Now III, IV, Lemma 3.1.3 and Theorem 3.1.2 imply what was re• quired. The assertion is proved.
2. Hahn-Banach theorem Theorem 2.1 (Hahn and Banach). Let L be a vector space. Let a function p : L --t [0,00) satisfy the conditions:
(2.1) p(u+v)
~p(u)
+p(v) for allu,v E L;
(2.2) p(au) = ap(u) for all u E L, a
~
O.
Let a linear functional fo : Lo --t IR be defined on a vector subspace Lo of the space L and satisfy the condition
(2.3) fo(u)
~
p(u) for all u E Lo.
Then there exists a linear functional f : L --t IR such that fiLa = 10 and -pC -u) ~ feu) ~ p(u) for all u E L. Proof. I. Use Lemma 1.8.2. Denote by A the set of all pairs (M, cp), where M 2 Lo is a vector subspace of the space L, cP : M --t IR is a linear functional, cplLo = 10 and cp(u) ~ p(u) for all u E M. The set A is partially ordered by the rule: (MI' cpd -< (M2' CP2), if MI ~ M2 and CP21 M1 = CPl' II. Show that if (MI' cpd E A and MI i= L then there exists an element (M2' CP2) >- (MI' cpd, M2 i= M I, of the set A. Since MI i= L, there exists a vector Wo E L \ MI' The set M2 = {u + AWo: u E M I , A E 1R} is a vector subspace of the space L. Since Wo E L \ M I , the representation of every element of the set M2 as a sum u + AWo is unique. Take an arbitrary c E lR. Consider the functional 'l/Jc(u + AWo) = CPI(U) + AC. on the space M2· Our nearest aim is to find a value of c such that the pair (M2' 'l/Jc) belongs to A. For WI, W2 E MI we have CPI(wd
+ CPI(W2) = CPI(WI + W2)
~ P(WI
+ W2)
+ (W2 + wo)) wo) + P(W2 + wo).
= P((WI - wo) ~
P(WI -
Therefore CPI(wd -P(WI -wo) ~ P(W2+ WO) -CPI(W2)' Thus the numbers = inf{p(w + wo) - CPI(W) : wE Md are related by the inequality ql ~ q2' ql
= SUp{CPI(W) - pew - wo): wE Md and q2
Weak topology on the space Ll and derivation of convergent sequences. 167
Take now as C an arbitrary number of the segment [ql, q2]. For every vector u + AWo E M2 with u E Ml and A =J 0 we have
For A > 0 the right hand inequality implies:
CA ~ p(u
+ AWo) -
CPl(U),
'l/Jc(u
+ AWo) =
CPl(U)
+ AC ~ p(u + AWo)·
For A < 0, when we multiply the left-hand inequality by -A, we obtain: CPl (u) - p(u + AWo) ~ -AC,
'l/Jc(u + AWo) = CPl (u)
+ AC ~ p(u + AWo)·
For A = 0
Thus (M2' 'l/Jc) E A. III. By Lemma 1.8.2 there exists a maximal linearly ordered (by the relation -< of I) set B S;; A. Its maximality means that the set B is not a proper subset of an other linearly ordered subset of the set A. Denote by M(3 the first element of the pair (3 = (M, cp) E B (i.e., the subspace M S;; L). Denote by CP(3 the second one (i.e., the functional cp). The set if = U{M(3 : (3 E B} is a vector subspace of the space L. We define uniquely the linear functional rp on if by the condition rpiM/3 = CP(3. Since functionals CP(3 satisfy (2.3), this condition is satisfied by the functional rp too. Thus (if, rp) E A. The pair (if, rp) is not less than every element of the set B (in the sense of the order of I). By virtue of the maximality of the set B the pair (if, rp) belongs to B. If if =J L then by II there exists an element of the set B following (ii, rp). This is impossible by the definition of (if, rp). Thus if = L. For U E if = L we have rp(u) = -rp( -u) ~ -p( -u). The theorem is proved. • Remark 2.1. The norm II . II satisfies (2.1 - 2). For every m > 0 the /unction p(u) = mllull satisfies (2.1 - 2). By Remark 2.1 Theorem 2.1 and Assertion 1.1 imply Theorem 2.2. Let Lo be a vector subspace of a normed space L. Let fo : Lo - t IR be a continuous linear functional. Then there exists a contin• uous linear functional f : L - t IR extending the functional fo. Theorem 2.3. Let L be a vector space. Let a function p : L - t [0,(0) satisfy conditions (2.1- 2) and Uo E L. Then there exists a linear functional f : L - t IR such that
168
CHAPTER 5
and -p( -u)
~
f(u)
~
p(u)
for every vector u E L. Proof. Define on M = {Auo : ). E JR} an (obviously, continuous) functional f 0 ().uo) = ).p( uo). Theorem 2.1 implies the existence of its extension f to L. The extension f satisfies the estimates -p( -u) ~ f(u) ~ p(u). The theorem is proved. • When we take the norm as the function p in Theorem 2.3, we obtain, in particular (with the notation of Remark 1.1), that for every vector u E L there exists a linear functional f E L* such that f(u) :f. O. This implies: Assertion 2.1. If u ELand u :f. 0, then (u) :f. O. • Moreover, the functional f mentioned in the proof of Theorem 2.3 satisfies conditions
f(uo) = Iluo II,
(2.4) (2.5)
If(u)1 ~
It follows from (2.4-5) that
Ilull for of all u IIfll
E L.
= 1. By (2.4)
Therefore By (1.1) this means that
11(uo)11
=
Iluoll·
Thus the mapping : L ---t (L*)* keeps the norm. So it is isometric. So the mapping . : L ---t (L*):, which coincides with as the mapping of the sets, is continuous. The bijection . gives the possibility of defining a new topology on the set L (denote the corresponding topological vector space by L.):
a set U ~ L. is open in the space L. if and only if the set .(U) is open in the space 8(L) ~ (L*):. The new introduced topology on the vector space L is called the weak topology of the normed space L. By virtue of Remark 1.1 and the isometry of the mapping the identity mapping L ---t L8 is continuous.
Weak topology on the space £1 and derivation of convergent sequences. 169 3. Convex sets and linear functionals Let A :3 0 be a convex open subset of a normed space L. Then G"O ~ A for some c > O. Therefore: for every vector U E L there exists a number A > 0 such that A-lU E A (it is sufficient to take A = ~, see E Figure 5.1).
AA
In other words the set AA = {.Aw: w E A} contains the vector u. So the set Figure 5.1 Su = {.A: A ~ 0, u E AA} is nonempty. The Minkowski functional of the set A is defined by the formula JLA(U) = inf suo By the remarks just made 0 ~ JLA(U) < 00. Assertion 3.1. Under the above assumptions: 1) JLA(U + v) ~ JLA(U) + JLA(V) for all u, VEL; 2) JLA(Au) = AJLA(U) for all U ELand A ~ 0; 3) A = {u: U E L, JLA(U) < 1}. Proof. I. Take arbitrary numbers a > JLA(U) and b > JLA(V). Then a- 1 u, b- 1 v E A. By virtue of the convexity of the set A the vector u +V _ -a- (-1) a U a+b a+b
b (b- 1 v ) +--
a+b
belongs to A. Therefore a + bE su+v. This means that a + b ~ JLA(U + v). By virtue of the arbitrariness in the choice of a and b this implies 1). II. The property 2) follows trivially from the obvious equality SAU = As u . III. The continuity of the operation of the multiplication of a vector by a number and of the openness of the set A implies 3). The assertion is • proved. Lemma 3.1. Let f : L ~ ~ be a continuous linear functional and f t= O. Let A be an open subset of the space L. Then the set f(A) is open in R Proof. Take an arbitrary x E f(A). Let U E A and f(u) = x. The condition f t= 0 implies the existence of a vector v satisfying the inequality f(v) > o. The continuity of the operations of the addition of vectors and of the multiplication of a vector by a number, the openness of the set A imply the existence of a number c > 0 such that U + .xv E A for
170
CHAPTER 5
all ). E (-e, e). It remains to notice that
f(A) ;;2 {f(u
+ ).v):
). E (-e, e)} = (x - ef(v), x
+ ef(v».
The lemma is proved. • Theorem 3.1. Let A and B be nonempty convex disjoint subsets of a normed space L. Let the set A be open. Then there exist a continuous linear functional f : L ---t lR and a number 'Y E lR such that
(3.1) f(u) < 'Y :::; f(v) for all u E A and v E B (Figure 5.2).
Proof. Fix an arbitrary couple of vectors Ua E A and Va E B. The set
+ Va Ua + Va
C= A- B = {u U
= u{ {u u E
A}:
B
Ua
V -
E A,
[
V
E B}
V -
Ua
V
A
:
+ Va:
E B}
is open in the space L and contains O. So
Figure 5.2
(3.2) for some e > O. The convexity of the sets A and B implies the convexity of the set C. Since AnB = 0 the vector Wa = Va -Ua does not belong to C. Therefore (3.3) For U E L put p( u) = ing the condition
(3.4)
::/::;>. Construct a functional f : L
---t
lR satisfy-
-JLe( -u) ~ -p( -u) ~ f(u) ~ p(u) ~ JLc{u) for all u E L
according to Theorem 2.3. By (3.2)
Weak topology on the space L1 and derivation of convergent sequences. 171 for every vector u E L. Therefore u E latter estimate and (3.4) imply: f(u) ~ c- 1 11ull
c 1 11ull· C,
f (u - v
Therefore f(u) - f(v)
(3.5)
f(u)
<
+ wo)
+ f(wo) = f(v)
~
f.
/La( u - v
f(u) - f(v)
for all
u EA
c 1 11ull. The
u E L.
for all
This means the continuity of the functional For arbitrary u E A and v E B we have
/La(u) ~
+ wo) < 1. + 1 < 1, and
v EB
So'Y < 00 for 'Y = sup{f(u): u E A}. The openness of the set A and Lemma 3.1 imply the openness of the set f(A). Therefore'Y > f(u) for every vector u E A. This observation and (3.5) imply the fulfilment of (3.1). The theorem is proved. • Theorem 3.2. Let A and B be nonempty convex disjoint closed subsets of a normed space L. Let the set A be compact. Then there exist a continuous linear functional f : L ---* ~ and numbers p < q satisfying the condition
(3.6)
f(u)
f(v) for all u E A and v E B.
Proof. By Lemma 2.3.8 the distance c = p(A, B) is positive. By Lemma 2.4.3 the set DcA is convex. By the choice of c we have DcAnB = 0. When we apply Theorem 3.1 to the couple consisting of the sets DcA and B, we obtain the existence of a continuous linear functional f : L ---* ~ and a number'Y satisfying (3.6). The set f(A) is compact. Therefore the number p = sup f(A) belongs to f(A). Hence p < 'Y. It remains to put q = 'Y. The theorem is proved. • 4. Weakly convergent sequences We say that a sequence a = {Uk: k = 1,2, ... } of elements of a normed space L converges weakly to u E L if the sequence a converges to u with respect to weak topology of the space L. By virtue of the definition of the weak topology a sequence {Uk : k = 1,2, ... } ~ L converges weakly to u E L if and only if f(uk) ---* f(u) for every functional f E L * . Remark 4.1. If a sequence a = {Uk: k = 1,2, ... } ~ L converges weakly to u E L, then u E cc(a). To prove it, consider a pair consisting of the one point set A = {u} and the closed convex set B = cc(a). If we
172
CHAPTER 5
assume that our assertion is false then An B = 0. By Theorem 3.2 there exists a continuous linear functional f : L - t IR such that f(u) < inf f(B). Then f(u) # lim f(ud (~f(B)). k->oo
This contradicts the initial assumption. Remark 4.2. Under the assumptions of Remark 4.1 we can fix positive integers kl < k2 < ... in such a manner that the vector u is a limit (with respect to the norm of the space L) of the sequence of vectors Vi E c( {Uk., ... , Uk.+l-d). The mentioned vectors may be pointed sequentially. Since u E cc(a) = [c(a)], Remark 3.1 implies the existence of a vector VI E c(a) n 0 1 u. For some kl we have VI E c( {Ul, . .. , Ukl-l}). So we have done the first step of our inductive construction. Assume now that the vectors VI, ... , Vi and the numbers kl' ... ' k i are fixed. Since u E CC({Uk" Uk.+!, ... }), as in the first step there exists a vector 'liHI E c( {Uk., Uk.+!, ... }) n O2 -, U. Moreover, Vi+! E c( {Uk., ... , Uk.+l- 1 }) for some kHI > k i . Since the sequence {Vi: i = 1,2, ... } satisfies the condition Vi E O2 -.+1 U, we have the convergence Vi - t U with respect to the norm. Assertion 4.1. Let {Uk: k = 1,2, ... } ~ Land
(4.1) for every continuous linear functional f : L {If(udl: k = 1,2, ... } be bounded.
Then sup{llukll: k Proof. Let
-t
IR the sequence
= 1,2, ... } < 00.
Mi=n{{f: fEL*, If(udl=:;;i}: k=1,2, ... }. 1. Because of (4.1) L*=U{Mi: i=1,2, ... }. The inequality in the definition of the set Mi is non-strong. Therefore the set Mi is closed in the space L *. Assertion 1.3 and Baire theorem 1.6.7 imply that for some j = 1,2, ... the interior (Mj ) of the set M j is nonempty. Therefore for some point g E M j there exists a number E E (0,1) such that
Oog
~
Mj
.
If now h E L* and Ilhll < k = 1,2, ... we have:
E
then g - h E M j
.
Therefore for every
II. By Theorem 2.3 for every k = 1,2, ... we can fix a functional fk E L* in a manner that fk(Uk) = IIUkll and IIfkll =:;; 1.
Weak topology on the space L1 and derivation of convergent sequences. 173 By I (with h =
~fk)
for every k = 1,2, .... The assertion is proved. Assertion 4.2. Let a sequence {Uk: k = 1,2, ... } to a vector u E L. Then
~
• L converge weakly
Proof. Assume in addition the existence of a limit lim Iluk II (if the k ..... oo limit does not exist we choose a corresponding subsequence). Consider a continuous linear functional f E L* such that f(u) = Ilull and Ilfll ::;;; 1. Such a functional exists by Theorem 2.3 (with the norm 11·11 as the function p). Then The assertion is proved. • Theorem 4.1. A sequence {Uk: k = 1,2, ... } ~ L converges weakly to a vector u E L if and only if the following two conditions hold at once: 1) sup{lIukll: k = 1,2, ... } < 00; 2) there exists a dense subset M of the space L * such that lim f(Uk) = f(u) for every functional f E M. k-+oo
Proof. The necessity of condition 1) follows from Assertion 4.1. The necessity of condition 2) is obvious. Let us prove the sufficiency of these conditions. Let f E L, m = sup{ Iluk II: k = 1,2, ... } and € > O. Consider an arbitrary functional 9 E Oef nM. We have:
If(ud - f(u)1 ::;;; If(Uk) - g(uk)1 + Ig(ud - g(u)1 ::;;; €llukll + Ig(ud - g(u)1 + €lIull, lim sup If(Uk) - f(u)1 ~ 2€m.
+ Ig(u)
- f(u)1
k ..... oo
By virtue of the arbitrariness in the choice of € > 0 the last estimate implies the equality lim sup If(Uk)- f(u)1 = O. Therefore lim f(Uk) = f(u). k-+oo
The theorem is proved.
k ..... oo
•
5. Spaces Ll and Loo Let n = 1,2, ... and -00 < a < b < 00. The space Ll = Ll (la, b]) mentioned in the title of the section consists of Lebesgue integrable functions
174
CHAPTER 5
I : [a, b] ~ ]R.n; moreover, if two such functions coincide almost everywhere we mean that they represent the same element of the space L 1 . The Lebesgue integrability of the function I and remarks of §4.7 imply the integrability of coordinate functions Ii, i = 1, ... ,n, of the function I. This implies the integrability of the function g(x) = max{IIl(x)I,···, IIn(x)I}. The inequality (2.2.1) and Lemma 4.7.6 imply that: (5.1)
the function 11111 Lebesgue integrable.
Vice versa, a measurable function I : [a, b] ~ ]R. satisfying (5.1) belongs to the space Ll by Lebesgue theorem 4.7.4. The structure of a vector space on the set Ll may be defined in a natural way and the function b
11111 =
J
III(t)1I dt
a
is a norm on the space L 1 • Introduce in the consideration the function 1111100 = inf{m : m ~ 0, /-l({t: t E [a,b], II(t)1 > m}) = O}. For some part of elements I of the space Ll 1111100 = 00. The set Loo = {f: I ELI, 1111100 < oo} is a vector subspace of the space L 1 . The function 1111100 is a norm on Loo. This norm is defined independently on the main norm II I II of the space L 1 • The topology generated by the norm 1111100 need not coincide with the topology generated on Loo by the norm 11111. On the other hand, (5.2)
11111 ~ (b - a)IIIlIoo.
Therefore we obtain the continuity of the identity mapping Loo ~ L 1 • Here we consider the space Loo with the norm 11·1100 and we consider the space Ll with the norm II . II ( (5.2) implies the satisfaction of Lipschitz condition). Theorem 5.1. The space Ll is complete. Proof. Let Ib 12, ... be a fundamental sequence of elements of the space L 1 • Fix indices i(l) < i(2) < ... such that and IIIi(k) - Ijll < 2- k for j ~ i(k). Pass to the subsequence CPk = Ii(k)l k = 1,2, .... In particular,
(5.3) Because of (5.3) the partial sums IIcp2 - CPIII + IIcp3 -CP211 + ... + Ilcpk -CPk-III of the series IIcp2 - cpIiI + IIcp3 - CP211 + ... are bounded in totality by the number ~ + ~ + ... = l. By Theorem 4.7.2 the series (5.4)
Weak topology on the space Ll and derivation of convergent sequences. 175 converges almost everywhere on the segment [a, b] to a (non-negative) function 1jJ(x). At every point of convergence of the series (5.4) the series
(x) = (CP2(X) - CPl(X))
(5.5)
+ (CP3(X)
- CP2(X))
+ ...
converges too; moreover 11(x)11 ~ 1jJ(x). Theorem 4.7.4 implies the Lebesgue integrability of the function (x). Let f(x) = CPl(X)+(x). For m = 1,2, ... put l(m) = max{k: k = 1,2, ... , i(k) ~ m}. Then
II -
IIf - fmll ~
(5.6)
CPI{m) II
+ Ilfi{l{m)) - fmll·
By virtue of the estimates
Ilf - CPkll = II + CP1 - CPkll = II(CPk+l ~ 2- k + 2- k- 1 + ... = 2- k+l
CPk) --t
--t
- cpk+d
+ .. ·11
0,
Ilfi{l{m)) - fmll ~ 2- i {l{m)) and (5.6) we have the convergence fm The theorem is proved.
+ (CPk+2 --t
0,
f in the metric of the space L 1 • •
6. Common representation of a linear functional on the space L1 Let
-00 < a < b < 00, n = 1,2, .... Let cP E Loo and e= Ilcplloo. Then b
(6.1)
J
b
(u(s), cp(s)) ds
a
~
JlIu(s)ll· Ilcp(s)11
b
ds
J
~ e lIu(s)11
a
ds =
ellull
a
for every function u ELI, where (.,.) in the first integral denotes the scalar product. Thus the formula
J b
f(u) =
(u(s), cp(s)) ds
a
defines a continuous linear functional on the space L 1 • Let us show that really this is a common representation of a continuous linear functional on the space L 1 • For x E [a, b] denote by ax the characteristic function of the subset [a, x] of the segment [a, b]: _ {
ax ( t ) -
1
o
if if
t E [a, xj, t E (x,bj.
CHAPTER 5
176
Remark 6.1. Let e be an arbitrary unit vector of the space JRn. Let the mapping [a, b] - Ll associate to a number x E [a, b] the function axe. The mapping is continuous. It is even isometric: b
Ilaxe - ayell =
y
J
J
a
x
lax(t) - ay(t)1 dt =
1 dt = Ix - YI·
Let f : Ll - JR be a continuous linear functional. Let e be an arbitrary unit vector of the space JRn. Remark 6.1 implies the continuity of the function rp(x) = f(axe). Let (6.2) , = {(ai, bi ): i of the segment [a, b],
=
1, ... , k} be a family of pairwise disjoint intervals a ~ al < bl ~ a2 < ... ~ ak < bk ~ b.
For i = 1, ... ,k put
O"(i) = {
1 -1
if if
f(abie) - f(aaie) ~ 0, f(abie) - f(aaie) < o.
Define the function f3 : [a, b] - JR,
f3(t) = {
o 0"( i)
U"
if t rJ. if t E (ai, bi ),
i = 1, ... , k.
The function f3 admits the representation
(in the sense of the space L l , so we do not distinguish functions coinciding almost everywhere). Therefore (6.3)
f(f3e) = O"(l)(f(ah e) - f(aal e)) + ... + l7{k)(f(ah e ) - f(aa. e )) = If(ab1e) - f(aale)1 + ... + If(ab. e ) - f(aake)l·
Assume now that (6.4) for i = 1,2, ... a family condition (6.2)
'i of intervals of the segment [a, b] satisfies
The sequence of corresponding functions f3i converges in the space Ll to the function [J == o. By virtue of the continuity of the functional f
Weak topology on the space L1 and derivation of convergent sequences. 177
f(f3ie) --t 0 as i --t 00. By (6.3) this means the absolute continuity of the function cpo Remark 6.2. Assume that U is the characteristic function of a set E ~ [a, b], e is an arbitrary unit vector ofthe space ~n, bi: i = 1,2, ... } is a sequence of families of intervals of the segment [a, bj satisfying (6.4), (6.5) p,( Gi
E \
E)
~ --t
Gi = Uri 0 as i --t
for every i = 1,2, ... , G1 ~ G 2 ~ G 3 ~ ••• and Ui is the characteristic function of the set Gi .
00,
The sequence of the functions Ui converges in the space L1 to the function u. By virtue of the continuity of the functional f
f( Ui e ) --t f( ue)
(6.6)
as
i
--t
00.
We have
f(Ui e ) = (f(ab 1 e ) - f(aal e » + ... + (f(ahe) - f(aak e » = (cp( ab 1 ) - cp( aal » + ... + (cp( abk) - cp( aak» (we keep the notation of the previous remark). Thus
J
f(Ui e ) =
cp'(s) ds.
Gi
By Theorem 4.7.3 and by (6.5-6)
J
J
E
E
cp'(s) ds =
cp'(s) ds
= .lim
'1.--+00
J
+ i~~
J
cp'(s) ds
Gi\E
cp'(s) ds = '1.--+00 .lim f(Ui e ) = f(ue).
Gi
Let e1, ... ,en be a basis of the space ~n, CPm(x) = f(axe m), cp = {CP1, ... , CPn}. By our remarks the function cp is absolutely continuous. Show that cp' E Loo and IIcp'lloo ~ Ilfll. Assume the opposite. Then the measure of the set M = {t: t E [a,bj, IIcp'(t) II > IIfll} is positive. Let A be the family of all closed balls S of the space ~n satisfying the conditions: all coordinates of the center of the ball S are rational; the radius of the ball S is rational; the ball S lies entirely outside the closed ball of the radius IIfll with the center at the origin of coordinates. The family A is countable: A = {Si: i = 1,2, ... }. We have M = U{Mi: i = 1,2, ... }, where Mi = {t: t E [a, bj, cp'(t) E Sd. By Theorem 4.1.2 JL(Mi ) =F 0 for some i = 1,2, ....
178
CHAPTER 5
Let e = {a1,'''' an} be the unit vector co-directed with the position vector of the center of the ball Si' Let u be the characteristic function of the set E = Mi. By Remark 6.2
f(ue) = ar!(ue1) + ...
+ anf(uen )
= a1 j cp~(s)ds + ... + an j cp~(s)ds E
E
(a1cp~ (s) + ... + ancp~(s))ds
= j E
= j(e,cp'(s))ds E
> Ilfll . Ji,(E) = Ilfll . Ilull = Ilfll ·lIuell· This contradicts Assertion 1.2. To complete the reasoning consider the continuous functional b
(6.7)
g(u) = f(u) - j(cp'(s),u(s)) ds. a
By Remark 6.2 g(u) = 0 for every product u of the characteristic function of an arbitrary measurable subset of the segment [a, b] by an arbitrary vector. By virtue of the linearity of the functional 9 we have g( v) = 0 for every sum v of such products. Such sums constitute an everywhere dense subset of the space L1 (see Example 4.7.6). Therefore the continuity of the functional 9 implies that 9 = O. In other words, b
f(u)
= j(cp'(s),u(s)) ds a
for every function u E L 1 • Theorem 6.1. A sequence {h: k = 1,2, ... } ~ L1 converges weakly to a function fELl if and only if the following two conditions hold at once: (6.8)
sup{llfkll: k = 1,2, ... }
< 00,
and (6.9)
I
D
D~Ia,bl·
h(s) ds
~
I
D
f(s) ds as k
~ 00
for every measurable set
Weak topology on the space Ll and derivation of convergent sequences. 179 Proof. Theorem 4.1 and our remarks imply that the convergence of the sequence Uk: k = 1,2, ... } to the function f is equipotent to the fulfilment at once of conditions (6.8) and b
(6.10)
b
j(h(s),u(s)) ds
~
a
jU(s),u(s)) ds a
as k ~ 00 for every sum u E Leo of products of characteristic functions by vectors. The equivalence of (6.9) and (6.10) is obvious (note that b
j h(s) ds = j h(s)XD(s) ds). D
a
•
The theorem is proved. 7. Space of measurable sets Let
-00
< a < b < 00.
Denote by Mo(a, b) the set of all measurable subset of the segment [a, b]. We regard sets A, B E Mo (a, b) equivalent if the measure of the set AAB = (A \ B) U (B \ A) is equal to zero. Notice that b
(7.1)
Jl(AAB)
= J IXA(S) - XB(s)1 ds = IIXA - XBII· a
Let the mapping <1>0 : Mo ~ Ll associate to a set its characteristic function. The images of sets A and B under the mapping <1>0 coincide if and only if the set A and B are equivalent. Therefore the mapping <1>0 generates the mapping <1> : M ~ Ll of the quotient set M of the set Mo with respect to the above equivalentness. Moreover (7.1) implies that the fUnction p(A, B) = Jl(AAB) is a metric on the set M and the mapping <1> turns out to be an isometry. Assertion 7.1. The metric space (M, p) is complete. Proof. Let a = {Ak: k = 1,2, ... } be an arbitrary fundamental sequence of elements of the space M. It is sufficient to prove the existence of a limit for a subsequence of the sequence a. Thus without loss of generality We can assume in addition that p(Ak' A,) < 2- k for every I > k. Let Bk = U{Ai: i = k, k + 1, ... }.
CHAPTER 5
180
peAk. B k ) ~ jl(U{Ai+1 \ Ai: i = k, k ~
2- k
+ 1, ... })
+ 2- k - 1 + ... = 2- k +l.
Therefore it is sufficient to prove the existence of a limit of the sequence (3 = {B k : k = 1,2, ... }. This is obvious: the set n(3 is the limit of the sequence (3. The assertion is proved. •
8. Weak compactness in the space L1 Keep the notation of the previous sections. Our aim in this section is to understand, when (8.1) every subsequence of a (the) sequence Uk: k = 1,2, ... } ~ L1 has a subsequence converging weakly to a function fELl' A sequence Uk : k = 1,2, ... } satisfying condition (8.1) is called relatively weakly compact (see §1.6). By Theorem 6.1 the fulfilment of condition (6.8) is necessary for the fulfilment of condition (8.1). In order to point a second necessary condition
consider following situation: (8.2) {Mk : k = 1,2, ... } is a sequence of subsets of the segment [a, b] and jl(Mk ) - t 0 as k - t 00.
k
Lemma 8.1. Let n = 1, condition (8.2) hold and the sequence converge weakly to the function f == O. Then
Uk
= 1,2, ... }
(8.3)
J
h(s) ds
-t
0 as k
-t
00.
Mk
Proof. 1. With the notation of §7 the set S =
J b
GJ(u)
=
f(s)u(s) ds
a
is continuous.
~ -t
L1 is complete. JR,
Weak topology on the space L1 and derivation of convergent sequences. 181 For
f
E
Loo this is obvious:
100f(u) - O!f(v)1
~
b
J
If(s)I·lu(s) - v(s)1 ds
~ Ilflloo ·llu - vii·
a
In the general case fix a sequence Uk: k = 1,2, ... } to the function f in the space L 1 • Then for every function u E S
~
L oo , converging
b
100fk(u) - O!f(u)1
~
J
IJk(S) - f(s)l· u(s) ds
~ Ilh - fll·
a
So we have the uniform convergence O!h ~ O!f. This implies what was required. II. Take an arbitrary c > O. For l = 1,2, ... put
By virtue of the weak convergence of the sequence Uk: k = 1,2, ... } to the function f == 0 we have b
O!fk (u)
=
J
h(s)u(s) ds
~0
a
for every u E S. Thus S = U{S/: l = 1,2, ... }. The sets S/ are closed in the space S (the functions O! h are continuous and the inequality in the definition of S/ is non-strong). By Baire's theorem 1.6.7 for some m = 1,2, ... the set (Sm) is nonempty. Therefore there are A E Sm and 6 > 0 such that the 6-neighborhood U of the point A in the space S lies in the set Sm. If BE U and n = m, m + 1, ... , then
III. If with the notation of II C ~ A and p,(C) ~ 6, then
IlxA -
J
fn(s) ds
C
=
J
XA\clI ~
fn(s) ds -
A
8,
J A\C
fn(s) ds,
182
CHAPTER 5
Jf (
C
~ "'8
8) ds
n
C
C
+ -8
= -
4·
C
IV. With the notation of III put C+ = {s: C _ = {s: sEC, f n(s) < O}. When we apply III to these sets we obtain
J
Ifn(s)1 ds
C
J
=
SEC, fn(s)
J
fn(s) ds -
fn(s) ds
~
O} and
~ ~ + ~ = ~.
c_
C+
V. If with the notations of II C ~ S \ A and /-l(C) ~ 15, then IlxA - XA\cll ~ 15,
J
J
fn(s) ds =
J
fn(s) ds -
fn(s) ds,
AuC
C
A
J
f n ( s) ds ~
C
C
C
8 + 8 = "4.
C
Now when we repeat the reasoning IV, we obtain
J
Ifn ( S ) I ds ~
C
C
"4 + "4
C
2·
=
C
VI. Let a set C
~
[a, b] be measurable, /-l( C)
J
Ifn(s)1 ds =
J
Ifn(s)1 ds
CnA
C
~
J
15 and n ~ m. Then
Ifn(s)1 ds
C\A
-C + -C =
'" 2
+
~
2
c.
The last estimate implies (8.3). The lemma is proved. Corollary. Let condition (8.2) hold. Let a sequence converge weakly to the function f == O. Then
J
fk(S) ds
---t
0 as
k
Uk:
• k = 1,2, ... }
---t 00.
Mk
To obtain the corollary notice that if g E (JRn)*, then the sequence {g(h): k = 1,2, ... } converges weakly to the function h == O. By Lemma 8.1
Weak topology on the space Ll and derivation of convergent sequences. 183
By virtue of the arbitrariness of 9 E (IR.n)* we have
• Assertion 8.1. Let conditions (8.1- 2) hold. Then condition (8.3) holds too.
Proof. Assume the opposite, i.e., for some
> 0 let the set
E
be infinite. By (8.1) the sequence Uk: k E A} has a subsequence Uk: k E Ad converging weakly to a function fELl. Since XMk --+ 0 almost everywhere, fXMk --+ 0 almost everywhere. Theorem 4.7.4 implies the convergence b
J
(S.4)
f(s) ds =
J
f(S)XMk(S) ds
--+
0
as
k
--+ 00
a
(we can take the function If I as a majorant). When we apply Lemma 8.1 to the sequence obtain
J
(S.5)
(Jk(S) - f(s)) ds
--+
0
Uk - f
as k E AI, k
k E
Ad,
we
--+ 00.
Mk By (S.4-f))
J
f k(s) ds =
Mk
as k
J
(Jk(s) - f (s )) ds
Mk
+
J
f (s) ds
--+
0
Mk
k --+ 00. So our assumption is false and the assertion is proved. We say that a sequence of functions 'Pk : [a, b] --+ ~n, k equi-absolutely continuous if: E AI,
=
• 1,2, ... , is
(8.6) for every E > 0 there exists 8 > 0 such that if a family {(ai, bi ) : i == 1, ... ,p} of pairwise disjoint intervals of the segment [a, b] satisfies
184
CHAPTER 5
the condition E{lbi - ail: i = 1, ... ,p} < 8 then E{II
-
i = 1, ... ,p}
Remark 8.1. If condition (8.6) holds then: (8.6') for every I:: > 0 there exists 8 > 0 such that if a family {(ai, bi ) : i = 1,2, ... } of pairwise disjoint intervals of the segment [a, bj satisfies the condition E{lbi - ail: i = 1,2, ... } < 8, then E{II
(8.6/1) for every J.L(M) < 8, then
I::
> 0 there exists a 8 > 0 such that if M
J
ds
C IR and
< I::
M
for every k
= 1,2, ....
To deduce (8.6/1) from Remark 8.1 fix 8 according to condition (8.6'), but for 1::/2 instead of 1::. By (8.6'): for every open set M k = 1,2, ...
~
[a, bJ of the measure < 8 and for every
J
< 1::/2.
M
Fix of a decreasing sequence of open sets
Weak topology on the space L1 and derivation of convergent sequences. 185 such that
Then for every k
f cp~(t)dt f cp~(t)dt + l~~ f i~~ f cp~(t)dt
M
M
cp~(t)dt
Mi\M
=
Mi
c ~-
[a, b]
--t
IRn , k = 1,2, ... ,
be equi-absolutely continuous and converge uniformly on the segment [a, b] to a function CPo. Then the function CPo is absolutely continuous. Proof. Take an arbitrary rJ > O. Fix 8 > 0 according to (8.6) for C = rJ/2. Let 'Y = {(ai,b i ): i = 1, ... ,p} be a family of pairwise disjoint intervals of the segment [a, b] with d(r) < 8. When we pass to the limit as k --t 00 in the last inequality of (8.6), we obtain: E{IICPo(bi ) - cpo(ai)11 : i = 1, ... ,p} ~ rJ/2 < rJ. This is what was required. The lemma is proved. • Lemma 8.3. Let t
CPi (t) =
J fi (s)
ds for
a
i == 0,1,2, ... and the sequence {CPi: i = 1,2, ... } converge uniformly to a /unction CPo.
Let conditions (6.8), (8.6) hold. Then
{8.8} condition {8.3} holds for every sequence of measurable sets {Mk k == 1,2, ... } satisfying {8.2}. and the sequence a converges weakly to the function 10 = cp~. Proof. By Theorem 6.1 and Assertion 8.1 it is sufficient to prove the fulfilment of {6.9}.
CHAPTER 5
186
1. For D = (c, d) ~ [a, bj (in (6.9) ) the convergence in (6.9) takes place by virtue of the convergence d
/ h(s) ds = 'Pk(d) - 'Pk(C)
~ 'Po(d)
d
- 'Po(c) = /10(s) ds.
c
c
II. By I in the case, when the set D may be represented as the union of finite number of intervals, the corresponding convergence takes place too. III. Let D be an arbitrary open subset of the segment [a, bj, 'Y = {G i : i = 1,2, ... } be the family of (all) its connected components. Take an arbitrary c > o. Lemma 8.2 implies the equi-absolute continuity of the sequence {'Pi: i = 0,1,2, ... }. Remarks 8.1-2 imply that for 'f/ = ~ (8.9) there exists 0
J.L(M) < 0, then
> 0 such that if the set M ~ [a, bj is measurable and every i = 0,1,2, ....
II J 1i(S) dsll < 1J for M
(This implies the fulfilment of (8.8).) Since J.L(D) = J.L(U'Y) ::;; b - a, there exists an index j = 1,2, ... such that J.L(U{ Gj , Gj+I, ... }) < o. Let H = G I U··· U G j - 1 . By II there exists an index l = 1,2, ... such that
/1k(S) ds - /10(s) ds < H
Thus for k
~
i
for
k
~ l.
H
l
/ 1k(S) ds - / 10(s) ds D
D
( /1k(S) ds H
+ /
1k(S) dS) - (/10(S) ds
D\H
: ; J1k(S) ds - J10(s) ds H
H
H
+
+ /
10(s) dS)
D\H
J 10(s) D\H
ds
Weak topology on the space Ll and derivation of convergent sequences. 187 By virtue of the arbitrariness in the choice of c this means the convergence
1 h{s) ds
-+
D
110{s) ds. D
IV. Let D be an arbitrary measurable subset of the segment [a, bJ. Take an arbitrary c > O. Fix 8 according to (8.9). By virtue of the measurability of the set D there exists a open subset H 2 D of the segment [a, bJ such that J.L{H \ D) < 8. By III there exists an index l = 1,2, ... such that
11k{s) ds - 1 10 {s) ds < H
Thus for k
~
i
for
k
~ l.
H
l
11k{S) ds - 1 10 {s) ds D
D
1
( 1 h{s) ds H
~
10{s) dS) - (110{s) ds -
H\D
1
1k{S) ds - 110{s) ds
H
c ~ ...., 3
H
+
H
c
1
10{s) dS)
H\D
1 10{s) ds H\D
c
+ -3 + -3 = c .
By virtue of the arbitrariness in the choice of c this means the convergence
1 1k{S) ds D
-+
110{s) ds. D
The lemma is proved. • Remark 8.3. Condition (8.8) implies that for every 'TJ > 0 condition (8.9) holds. If we take in (8.9) an open set (it falls into the union of the family'Y of its connected components) we obtain the fulfilment of (8.6) (see the notation in (8.7)).
CHAPTER 5
188
Assertion 8.2. Let a = Uk: k = 1,2, ... } ~ L 1 • Let conditions (6.8) and (8.9) hold. Then the sequence a satisfies condition (8.1). Proof. Let t
J
fi(s) ds for i = 1,2, ....
a
For arbitrary 'f/ > 0 fix 8 > 0 according to (8.9), see Remark 8.3. In particular, if c < d are points of the segment [a, b] and d - c < 8, then lI 0 according to (8.9) for 'f/ = 1. Choose points
to
= a < tl < ... < tj = b
in a manner that ti - t i- l < 8 for all i = 1, ... ,j. Take an arbitrary k = 1,2, .... For i = 1, ... ,j put Mi = {t: t E [ti-l, til, fk{t) ~ O} and Ni = [ti-l, til \ Mi· Then
11M =
!
1/.(t)l dt =
t (j
I.(t) dt -
!
I,(t) dt) :;; 2j.
The last estimate does not depend on k. The theorem is proved. • We have obtained a working (that is, applicable in real situations) test of the weak relative compactness in the space L 1 . Remark 8.4. Let a sequence of functions Uk: k = 1,2, ... } satisfy (8.8). Then the sequence of the functions {Ihl: k = 1,2, ... } satisfies (8.8) too. To prove it with the notation of (8.2) put M; = {t: t E Mk, fi{t) ~ O} and Mk* = Mk \ M;. Then
J
Ifk(S)1 ds =
~
J
fk(S) ds -
~
J
fk(S) ds.
~.
Both terms in the right hand side of the equality tend to zero. This gives what was required.
Weak topology on the space L1 and derivation of convergent sequences. 189
9. Derivation of a convergent equi-absolutely continuous sequence of functions Consider the following situation: (9.1)
-00
< a < b < 00, n
= 1,2, ... ;
(9.2) a sequence of functions CPk : [a, bj absolutely continuous;
--t
]Rn,
k
1,2, ... , is equi-
(9.3) the sequence {
Our aim is to prove the fulfilment of the condition
(9.4)
t
E
cp~(t) E
n{ cc( {cp~(t): i = k, k + 1, ... }): k = 1,2, ... } for almost all
[a,bj.
Remark 9.1. The fulfilment of conditions (9.1-3) implies that the function CPo is also absolutely continuous, see Lemma 8.2. Remark 9.2. The fulfilment of the conditions (9.1-3) implies that:
(9.5) for every c > 0 there exists a 8 > 0 such that if a set M ~ [a, bj is measurable and J.t(M) < 8, then II J cp~(t)dtll < c for every k = 1,2, ... , M
see Remark 8.2. Remark 9.3. By Remark 9.2 and Theorem 8.1 we have proved: Lemma 9.1. Let conditions (9.1 - 3) hold. Then the sequence of /unctions {
L1 •
•
Remark 9.4. By Lemma 9.1 and Assertion 8.2 we can assume in addition that the sequence {cp~: k = 1,2, ... } converges weakly to a function '¢. Since integral is a continuous functional, for every t E [a, bj we have t
CPo(t) - cpo(a) =
i~~(cpk(t) -
CPk(a)) =
i~~J cp~(s)ds = a
t
J 'Ij;(s)ds. a
Therefore cp~(t) = 'Ij;(t) for almost all t E [a, bj. Remark 9.5. By Remark 4.2 the function 'Ij; is the limit in L1 of convex combinations ~1' ~2"" of elements of the sequence {cp~ : k = 1, 2, ... }. So ~i(t) E cc({cp~(t):
k = 1,2, ... })
for almost all t E [a, bj and for every i = 1,2, ....
190
CHAPTER 5
By the reasoning in the proof of Theorem 5.1 we can choose a sequence i = 1,2, ... } in a manner that ~i(t) -+ 1f;(t) for almost all t E [a, b]. Therefore {~i:
(9.6)
cp~(t)
E cc( {cp~(t): i = 1,2, ... }) for almost all t E [a, b].
We can now shorten the sequence {CPk : k = 1,2, ... } and reject its first (k - 1) members. For the new sequence the analog of (9.6) remains true, namely cp~(t) E
cc( {cp~(t): i = k, k
+ 1, k + 2, ... })
for almost all t E [a, b].
.
Since this holds for every k = 1,2, ... we obtain: Assertion 9.1. Let conditions (9.1- 3) hold. Then condition (9.4) holds ~.
CHAPTER 6
BASIC PROPERTIES OF SOLUTION SPACES
Our aim in this chapter and in part of the next one is to point out a quite short list of properties of solution sets of ordinary differential equations which may be considered as the axiomatics for an appreciable part of the theory. In fact, the framework of the new theory will contain not only solution spaces of ordinary differential equation, but other objects too, although they are close to solution spaces with respect to their properties. Such an enlargement of the domain of action is not now essential, but the fact that new tools suit well the investigation of equations with different singularities in the right hand side, of corresponding differential inclusions, and of equations with the control, etc., turns out to be important. As axioms of our theory we take the simplest properties of solution sets, which are in fact constantly used in the theory of ordinary differential equations, although sometimes their statements, which are applicable to our investigation, look unusual in comparison with the corresponding conditions of the classical theory. The presence of some of these properties in solution spaces follows directly from the definition of a solution. For other properties we point out sufficient conditions. We do not want to finish the investigation in this chapter. Here our aim is quite narrow: we show that the axioms of our theory correspond to known properties of solutions of equations. Here we do not restrict ourselves to the simplest statements of sufficient conditions. As far as can be done, and with the use, in essence of the material of the previous chapters we prepare the ground for the latter account. We try to avoid here any complications in proofs and the introduction of cumbersome conditions which can make the initial grasp of the problem difficult.
1. Ordinary differential equations An equality of the type (1.1)
cI>1 (t, y(t), y'(t), ... ,y(m)(t)) = cI>2 (t, y(t), y'(t), ... , y(m)(t)) ,
or, in shortened notation, cI>l(t,y,y', ... ,y(m») = cI>2(t,y,y', ... ,ym), is called an ordinary differential equation. It relates values of the argument t
192
CHAPTER 6
and the corresponding values y(t), y'(t), ... , y(m)(t) of the unknown function y and its derivatives. The argument t takes real values and the function y (and its derivatives) take values in the space ]Rn (respectively, in the space en). The functions <1>1 and <1>2 are defined on corresponding subsets of the product ]R x ]Rn x ... x]Rn (respectively, ]R x en x ... X en) and take values in the space ]Rk or k . In fact, in what follows we consider equations of a more special type, namely of the type y' = <1>(t, y); however, this is often sufficient. For instance, when <1> and y in (1.1) are scalar functions then equation (1.1) may often be solved with respect to highest derivative of y. Our equation turns into an equation of the type
e
y (m) -_
Under the change Y1 = y, Yk changes into the system
=
y(k-1) for k
= 2, ... , m the last equation
y~ = Y2, y~ = Y3,
y~-l = Ym, y~ =
The new system may be written as a single vector equality y'
= <1>(t, y) with
<1>(t,X1,""Xm) = (X2,X3, ... ,Xm,
Often such transforms are possible for the vector functions <1> and y. Of course, here we can meet refinements related to the equivalence in such passages, but we will not digress to discuss it. The complex plane e may be considered as a vector space over R In this sense it is isomorphic to the space ]R2 (we associate to a complex number a + ib the vector {a, b} ). The space em may be also considered in an analogous way as a vector space over R In this meaning it may be identified with the space ]R2m. Because of the existence of the isomorphisms mentioned the consideration of an equation with a complex unknown function reduces to the real case, although in some situations the use of complex numbers makes finding solutions easier. 2. Solutions of differential inclusion y' E F(t, y)
Apart from ordinary differential equations we consider differential inclusions. Differential inclusions (== equations with multi-valued right hand
193
Basic properties of solution spaces.
sides) generalize differential equations. In the previous section we restricted our considerations to equations of a particular type. In this section we consider inclusions of a particular type also, namely, of the type mentioned in the title. Such inclusions are interesting because they both appear in applications and they are constantly used as tools in the investigation of differential equations. As in the case of an equation the complex case here in general reduces to the real one. So in what follows we study the real case only. Let F be a multi-valued mapping of an open subset U of the space IR x IRn into the space IRn. (Here we do not impose on the mapping F any conditions as the continuity or the measurability and we do not assume the nonemptiness of its values.) A function Z E G.(U) is called a solution of the differential inclusion (= solves the inclusion) y' E F(t, y) (or in the expanded form y'(t) E F(t, y(t)) ) if either the domain of the function z consists of one point only or the domain of the function z is nontrivial, the function z is generalized absolutely continuous, the approximate derivative z'(t) exists and belongs to the set F(t, z(t)) at almost every point t E 7r(z). In the case when the mapping F is single valued the differential inclusion y' E F(t, y) is usually written in the form of the differential equation y' = F(t, y), moreover the above definition of a solution of an inclusion turns into the definition of a solution of an equation. The solution set of a differential inclusion (equation) y' E F(t, y) will be denoted by D(F). For a subset M of the set U we denote D(F, M) = {z: z E D(F), Gr(z) ~ M}. The set D(F) lies in the topological space G.(U). With respect to the induced topology the set D(F) is also called the solution space of the inclusion y' E F(t, y). Notice some properties of solution sets following immediately from our definitions. Let Z = D(F). Then: (1) if z E Z and a segment I lies in 7r(z), then
Z2
ZII
E Zj
(2) if the domains of functions Zl, Z2 E Z intersect and the functions coincide on the set 7r(zd n 7r(Z2), then the function for t E 7r(zd for t E 7r(Z2),
Zl,
194
CHAPTER 6
7f(z) = 7f(zd U 7f(Z2), belongs to Z (see Lemma 4.9.1). These conditions are satisfied not only by solution sets of equations and inclusions. They are satisfied, for instance, by the space Cs(U), which is not the solution space of an equation or an inclusion, because it contains functions that is not generalized absolutely continuous. In order to deal freely with such sets let us introduce the following notation. Let M be a subset of the product lR x X of the real line lR and a topological space X. Denote by R(M) (respectively, by Ri(M)) the set of all subset Z of the space Cs(M) satisfying conditions (1) and (2) (respectively, condition (1)). Evidently conditions (1) and (2) are insufficient to develop a very valuable theory. Below we introduce other conditions also, which are satisfied (or may be satisfied) by the solution sets of equations and inclusions. We denote them for the convenience of references by small letters of the latin alphabet, for instance, (c), (e), (k) and so on. If * is an arbitrary set of such letters (conditions), then R~i) (M) denotes the set of all elements of the set R( i) (M) satisfying all conditions from *. We meet two first such conditions in the following lemma and its corollary. Lemma 2.1. Let F : U --t lRn be a multi-valued mapping and Z = D(F). Then
(q) the set Z consists of generalized absolutely continuous functions and if a function z E Cs(U) is generalized absolutely continuous, M is a closed subset of 7f(z), J1.(M) = 0 and ZII E Z for every segment I ~ 7f(z) \ M, then z E Z. Proof. The first part of (q), i.e., the assertion about the generalized absolute continuity, follows immediately from the definitions. Check the fulfilment of the second part of (q). By virtue of the openness of the set 7f(z) \ M in 7f(z) (and the regularity of a segment, see Lemma 2.3.9) for every point t of the set 7f(z) \ M there exists a segment It ~ 7f(z) \ M, which contains the point t in its interior (with respect to the segment 7f( z)). The family "( = {(It): t E 7f( z) \ M} is an open cover of the set 7f(z) \ M. By Theorem 1.6.10 the family "( contains an (at most) countable cover "(1 = {(It;}: i = 1,2, ... } of the set 7f(z) \ M. By our assumptions ZIIt. E D(F) for every i = 1,2, .... By the definition of a solution this me~ns the existence of a set Mi ~ It. (may be, empty) of measure zero such that at every point s E It. \ Mi the approximate derivative z' (s) exists and belongs to F(s,z(s)). By Theorem 4.1.2 the measure of the set M* = M U (U{Mi: i = 1,2, ... }) is equal to zero. At every point s of the set 7f( z) \ M* the approximate derivative z' (s) exists and belongs to the set
Basic properties of solution spaces.
195
F(s, z(s)). This means that the function z solves the inclusion y' E F(t, y). The lemma is proved. • It is convenient to highlight a particular case of the property of solution spaces from Lemma 2.1. When we take as the set M of the lemma the set consisting of two endpoints of the segment 7r(z), we obtain: Corollary. Let F : U --t ~n be a multi-valued mapping, Z = D(F). Then (p) the set Z contains all functions from Cs(U) defined on one point sets and if z E Cs(U), 7r(z) = la, b], a < b, and ZII E Z for every segment I ~ (a, b), then z E Z.
Notice that the first part of (p) follows immediately from definitions. In order to prove the fulfilment of the second part of (p) with the help of Lemma 2.1, we need to show the generalized absolute continuity of the function z. By our assumptions and by the definition of a solution the function ZII is generalized absolutely continuous for every segment I ~ (a, b). This implies immediately what was required. •
3. The inequality lIy'(t)11 ~ cp(t) Let -00 ~ ao < bo ~ 00, U = (ao, bo) x ~n. Let a non-negative function cp be defined and be locally Lebesgue integrable on the interval (ao, bo). The inequality from the title of the section may be understood as the differential inclusion y' E G(t) with G(t) = 0,(0, cp(t)). By Lemma 4.7.6 all solutions of this inclusion are absolutely continuous. Let q, be an arbitrary primitive of the function cp on the interval (ao, bo). For Z ~ Cs(U) and M ~ U put ZM = {z: z E Z, Gr(z) ~ M} (Thus D(F, M) = (D(F))M). Lemma 3.1. Let Z = D(G). Then (s) for every compactum K ~ U the set ZK is equicontinuous.
Proof. The projection Kl ~ ~ of the compactum K in the first factor of the product ~ x ~n is compact (see Lemma 2.1.1 and Theorem 1.7.8). The points a = inf Kl > -00 and b = sup Kl < 00 belong to Kl (see Theorem 1.6.1) and [a, b] ~ (ao, bo). Take an arbitrary number c: > O. According to the condition in the definition of the absolute continuity (see §4.5) find a number 8 > 0 such that for every family 'Y = {(ai,l3i): i = 1, ... , k} satisfying condition (*,8, [a, b]) of §4.5
196
CHAPTER 6
If now z E D(G,K), a,f3 E 7r(z) and la - f31 < 8 (hence the family {( a, f3)} satisfies the condition (*,8, [a, b])) then {3
{3
J
Ilz(f3) - z(a)1I =
J
~
zl(t)dt
cp(t)dt =
~c
'"
'"
(see Lemma 4.7.6). The lemma is proved. • Lemma 3.2. For every segment I ~ (ao, bo) the set D(G) n 7r- 1 (I) is closed in the space C(I, IRn). Proof. Let a sequence {Zj : j = 1,2, ... } ~ D( G) n 7r- 1 (I) converge uniformly to a function z (E C (I, IR n ) ). By Lemma 5.8.2 the function z is absolutely continuous. If the segment I consists of one point, then our assertion is trivial. In the opposite case for every point s of a subset M ~ I of full measure the derivatives Zl (s) and
Ilz(s + 8) - z(s)1I =
~im
)->00
Ilzj(s + 8) - zj(s)11 = )->00 ~im
8+6
J
z;(t)dt
8
~
8+6
J
cp(t)dt = I
8
Therefore
Ilz(S+8l-z(s)1I ~ 1
When we pass to the limit as 8
Ilz'(S)11
~
---t
0, we obtain
I
•
So z E D(G). The lemma is proved. Lemma 3.3. Let Z = D( G). Then
(n) all functions of Z are generalized absolutely continuous and if functions Zj E Z, j = 1,2, ... , are defined on a segment I of the real line and the sequence {Zj: j = 1,2, ... } converges uniformly to a function z E Z, then for almost all tEl the derivatives Zl(t), zj(t), j = 1,2, ... , exist and (3.1)
Z' ( t) E
cc ( { zj (t) : j = 1, 2, ... }).
Proof. The proofreduces to referring to Assertion 5.9.1. Condition (9.2) is needed for this referring. Its fulfilment follows from the estimate
J
zj(t)dt
M
~
J
cp(t)dt
M
Basic properties of solution spaces.
197
•
(see also condition (9.5) ). 4. Some properties of solution spaces
Properties established in lemmas of the previous section are helpful in the investigation of solution spaces. They hold place in situations that are very remote from the hypotheses of the lemmas. So it is convenient to make these properties a subject of an independent investigation without the assumption of any direct relation with restrictions on right hand sides of equations and inclusions. In this section we make some simplest remarks about this topic. First of all we enlarge the list of conditions under consideration and introduce the condition:
(c) for every compact urn K
~
U the set ZK is compact (with respect to the topology of the space C s (U) ). We will impose it on spaces Z E R(U), where U is a locally compact subset of the product IR x X of the real line IR and of a metric space X. We will also consider the condition of closedness in C s (U) of spaces under consideration. Lemma 4.1. A set Z ~ ~ Cs(U) is closed in the space if and only if for every compactum K ~ U the set Z K is closed in .
Proof. Necessity of the stated condition follows from the closedness of the set Cs(K) (see the definitions of the Vietoris topology and of the space Cs(M» and from the representation ZK = Z n Cs(K). Sufficiency. Let Z E [Z]. Lemma 2.4.2 implies the existence of a compactum K ~ U such that Gr( z) ~ (K). The set V = C s( (K)) is a neighborhood of the point z in the space Cs(U). The set ZK is closed. If it does not contain z then the set W = (V n without elements of Z. But the existence of such a neighborhood contradicts the condition z E [Z] <1>. Thus z E Z K ~ Z. The lemma is proved. • Lemma 4.2. Let ~ Cs(U), Z E Rp(U). Let for every compactum K ~ U and for every segment I ~ IR the set ZK n 7r- 1 (1) n be closed in (== in C(I, X) n , see Theorem 3.5.3). Then the set Z n is closed in . Proof. By Lemma 4.1 it is sufficient to show the closedness of the set ZK n in the space for every compactum K ~ U. By virtue of the IIletrizability of the space Cs(U) it is sufficient to show that the limit z E of every (convergent in belongs to ZK (see Theorem 1.5.3). Let 7r(z) = [a, b]. If a = b, then Z E Z by the first part of condition (p). Let a < b. Take arbitrarye E (0, b;a).
198
CHAPTER 6
Since the mapping 7r is continuous, 7r(Zj) 2 [a + c, b - c], beginning with some j. By Theorem 3.5.4 (4.1) E ZKn7r- 1 ([a+c,b-cD. The closedness of the set ZK n 7r-l([a + c, b - cD n and (4.1) imply that ZI[a+c,b-c] E ZK. By conditions (1) and (p) of §2 Z E ZK. The lemma is proved. • Theorem 3.7.1 and Lemma 4.2 (with = Cs(U)) imply: Theorem 4.1. Let Z E Rps(U). Let for every compactum K ~ U and every segment I ~ IR the set ZK n 7r- 1 (I) be closed. Then the space Z We have:
Zjl[a+c,b-c]
satisfies condition (c).
•
When we compare Theorem 4.1 with Lemmas 3.1 and 3.2, we obtain: Corollary. Under the assumptions of §3 D(G) E Rc(U). • Let us introduce some new notation. For a set M ~ IRx X and points t E R,yEXputMt={x: xEX, (t,x)EM}andMY={x: xEIR, (x,y)E M}. For a multi-valued mapping F : M ---t B (B is an arbitrary set) denote: Ft(x) = F(t,x) and FY(x) = F(x,y) (so we define the mappings Ft : M t ---t Band FY : MY ---t B). Return now to equations and inclusions. For X = IR n and an open subset U of the product IR x IR n denote by Q*(U) the set of all multi-valued mappings F : U ---t IR n with closed convex values satisfying at every point (t, y) E U the condition
(4.2) F(t,y) 2 n{cc(Ft(Oy)): Oy is a neighborhood of the point y in the space Ut (~ IRn)}. By Theorem 2.5.2 if the set F(t, y) is nonempty, convex and compact, then (4.2) is equipotent to the upper semicontinuity of the mapping F t at y.
Lemma 4.3. A space Z ~ Cs(U) satisfies condition (n) if and only if
(n') all functions of Z are generalized absolutely continuous and if functions Zj E Z, j = 1,2, ... , are defined on a segment I, the sequence {Zj: j = 1,2, ... } converges uniformly to a function Z E Z, then for almost all tEl the derivatives z'(t), zj(t), j = 1,2, ... , exist and (4.3)
Z'(t) En{cc({zj(t): j=k,k+1,k+2, ... }): k=1,2, ... }.
Proof. Condition (4.3) is equipotent to the fulfilment of condition (3.1) for all sequences
{Zj: j
= k, k + 1, k + 2, ... },
k = 1,2, ... ,
Basic properties of solution spaces.
199
at once. This gives what was required. The lemma is proved. • Theorem 4.2. Let a space n 7[-1(1). Let a sequence {Zj : j = 1,2, ... } ~ D(F) n E Rc(U). Then Z E Rc(U). Proof. Let K be a compact subset of the set U. By virtue of the obvious representation ZK = Z n of solutions of this inequality satisfies condition (n). By Corollary of Theorem 4.1 it satisfies condition (c). By Theorem 4.2 the space D(f, Ud is closed in
CHAPTER 6
200
we say 'initial values' in the plural we have in mind that Yo and values of the unknown function belong to the space ~n and the full coordinatewise record of the vector equation y( to) = Yo consists of n equalities). The Cauchy problem for an inclusion y' E F(t, y) with the initial values y(t o) = Yo (without the mentioning of the segment I) means the problem of the finding of a solution Z E D(F) such that to E (-7r(z)) and z(t o) = Yo. In this section we consider the existence of a solution (== the solvability) of the Cauchy problem in the particular case when the right hand side of the inclusion is single valued (Le., we deal with an equation y' = f(t, y) ) and satisfies the additional conditions:
a) for every t E ~ the function ft is continuous and for every y E ~n the function fY is measurable; b) -00 < ao < bo < 00, functions
for every (t, Yl), (t, Y2) E U. By virtue of the openness of the set U in ~ x ~n for every point (to,yo) E U there exists c > 0 such that (to - 2c, to + 2c) X G2 ,yo ~ U. Hence [to - c, to + cl x [G,yol ~ U. The function fl[to-"to+']x[O,yo] satisfies condition (4.13.1). By Theorem 4.13.3 it satisfies (4.13.2) too. By Corollary of Theorem 4.13.1 if z E Cs(U) and Gr(z) ~ [to - c, to + cl x [G,yol, then the function 7jJ(t) = f(t, z(t)) is measurable. By virtue of the compactness of the set Gr(z) and the continuity of the mapping a(t) = (t, z(t)) (see Theorem 2.1.1) the domain of every function z E Cs(U) may be covered by a finite number of segments II"'" h in a manner that for every j = 1, ... ,k the graph of the function ZII; lies in a set of the mentioned type. Our remark and Theorem 4.2.3 imply the measurability of the function 7jJ(t) = f(t, z(t)) for every z E Cs(U). Lemma 4.7.6 implies the Lebesgue integrability of the function 7jJ. If a function z solves the Cauchy problem y' = f(t, V),
(5.1)
y(t o) = Yo
on a segment 13 to, then for every point tEl (5.2)
z(t) =
(Z(t o) +
1
z'(s)ds
~)
t
Yo
+
J
f(s, z(s))ds.
to
Vice versa, by virtue of the above remark the function in the last integral of (5.2) is Lebesgue integrable. Hence every function z satisfying (5.2) is
201
Basic properties of solution spaces.
absolutely continuous. The derivation of the both sides in (5.2) implies that the function z solves the equation y' = f(t,y). The substitution t = to implies that z(t o) = Yo, i.e., every function z satisfying (5.2) solves the Cauchy problem (5.1). Theorem 5.1. Let a function f : U ~ lRn satisfy conditions a) and b). Then for every (to, Yo) E U there exists c > 0 such that for every segment 1 ~ (to - c, to + c), 1 3 to, there exists a unique solution of the problem (5.1) on I. Proof. For i = 0,1 and t E lao, bol put t
J
=
'Pi(s)ds.
to
Since U is open in lR x lRn , (to - 20, to + 20) X 026Yo The functions
~
U for some 0 > o.
space C(I, [06YO]) is complete. By virtue of the continuity of the mapping Q : C(I, [06YO]) ~ ]Rn, Q(z) = z(t o), the set X = Q-1(yO) is closed in C(1, [06YO]). Hence it is complete too. Associate to an arbitrary function z E Cs(U), 7I"(z) = 1, the function Az, t
Az(t) = Yo
+
J
f(s, z(s))ds.
to
By Lemma 4.7.6 t
IIAz(t) -
Yoll
=
J
~
f(s, z(s))ds
to
t
J
'Po(s)ds < 0
to
for tEl. So Az E X. We have the mapping A : Cs(U) n 71"-1(1) ~ X. By above remarks, a function z E Cs(U), 7r(z) = I, solves the problem (5.1) if and only if Az = z. Thus solutions of the Cauchy problem (5.1) on the segment I are fixed points of the mapping Alx : X ~ X. By Lemma 4.7.6 t
IIAz2(t) - Az1(t)1I
~
=
t
J
J
to
to
f(S,Z2(S))ds -
f(S,Z1(S))ds
t
J
IIf(s,Z2(S)) - !(s,z1(s))lIds
to
~
t
J
'P1(s)lIz2(S) - z1(s)lIds
to
202
CHAPTER 6
1
~ IIz2 - zlll·lcpl(t) - CPl(to)1 ~ 2"IIZ2 -
zd·
for Zl, Z2 E X and tEl. Since this is true for every tEl,
1 IIAz2 - AZll1 ~ 2"llz2 - zlll· To complete the proof it remains to refer to Theorem 1.7.4. The theorem is proved. • We will use the words 'Cauchy problem' not only in connection with differential equations or inclusions, but also in connection with an arbitrary space Z E R(U). Here U lies in the product ~ x X of the real line ~ and a metric space X. The Cauchy problem for a space Z with initial values y(t o) = Yo (on the segment I :1 to) means the problem of the finding of a function Z E Z, (7r(z)) :1 to (respectively, 7r(z) = I), z(t o) = Yo. When here the solution space of a corresponding differential equation or inclusion appears as the space Z, the new notion coincides with the notion of the above Cauchy problem. Lemma 5.1. Let Z E R(U). Then the following conditions are equipotent: a) for every point (to, Yo) of the set U there exists a number c > 0 such that for every segment I ~ (to - c, to + c), I :1 to, there exists no more than one solution of the Cauchy problem Z E Z, z(t o) = Yo on the segment I; b) for every point (to, Yo) of the set U there are its neighborhood V ~ U and a number c > 0 such that for every segment I ~ (to - c, to + c), 1:1 to, there exists no more than one solution of the Cauchy problem Z E Zv, z(t o) = Yo on the segment I; c) if the domains of functions Zl, Z2 E Z coincide and Zl(t) = Z2(t) at a point t of their common domain, then Zl = Z2; d) if the domains of functions Zl, Z2 E Z intersect and Zl(t) = Z2(t) at a point t E 7r(zd n 7r(Z2), then ZIi1l"(ztln1l"(Z2) = z2111"(ztln1l"(z2)' Proof. a=> b. It is obvious: we put V = U. b=>c. Show that the set M = {t: t E 7r(Zl), Zl(t) = Z2(t)} is open in the segment 7r(zd. Take an arbitrary to E M. Find a neighborhood V ~ U of the point (to, Zl (to)) and a number c > 0 according to condition b. By virtue ofthe continuity of the functions a(t) = (t, Zl (t)) and ,B(t) = (t, Z2(t)) (see Theorem 2.1.1) there exists a number 8 E (0, c) such that (t, Zl (t)) E V and (t, Z2(t)) E V for every point t E (to - 28, to + 28) n7r(zd. The functions Zl and Z2 coincide on the segment [to - 8, to +8] n7r(Zl)' Hence they coincide on the open subset (to - 8, to + 8) n 7r(Zl) of the segment 7r(zd. In view of the arbitrariness of the point to E M this proves the openness of the set M in 7r(zd. By Lemma 1.7.4 the set M is closed in 7r(Zl)' The set M is nonempty. The connectedness of the segment (see Example 2.6.1) implies the equality M = 7r(zd. .
Basic properties of solution spaces.
203
c {:}d. It is obvious. c =>a. It is obvious. The lemma is proved. • For the convenience in the further account we will word the just established and discussed properties in the following way:
(e) for every point (t, y) of the set U the Cauchy problem Z E Z, z(t) = y has a solution (Le., there exists z E Z, for which t E (-rr(z)) and z(t) = y); (u) if the domains of the functions Zl, Z2 E Z coincide and at a point t of their common domain Zl(t) = Z2(t) then Zl = Z2 (Le., the functions coincide on the entire domain of definition). The following assertion is obvious: Lemma 5.2. Let Z ~ Cs(U). Assume that for every point x of the set U there exists its neighborhood Ox ~ U such that the space ZOx satisfies condition (e) (on the set Ox). Then the space Z satisfies condition (e) (on the entire set U). • Pay attention to two particularities offacts proved in Lemma 5.1. First, the condition of the uniqueness, which occurs in Theorem 5.1 and which is mentioned in the lemma as condition a), turns out to be equipotent to the simpler in the statement conditions c), (i.e., (u)), and d). Secondly, the fulfilment of the condition of uniqueness (in every version) locally (see b) ) implies it is fulfiled on the entire set U. Thus, in particular, as a direct consequence of Lemma 5.1 we obtain: Lemma 5.3. Let Z E R(U). Assume that for every point x of set U there exists its neighborhood Ox ~ U such that ZOx E Ru (Ox). Then • Z E Ru(U). In order to be rid of necessity of using the word 'set' too often in connection with the sets R.(U), where * denotes an arbitrary set of conditions under consideration (properties of solution spaces), sometimes we will also call them classes of solution spaces. Their elements need not necessarily be solution spaces of ordinary differential equations or inclusions but they possess many properties of real solution spaces. In considerations below the classes Rceu(U) ~ Rce(U) will be most important. We can expand to them many assertions describing properties of solution spaces of ordinary differential equations. The use of other conditions in our list will often be related to the proof of the belonging of a space to these classes. In particular, we mean the verification the fulfilment of condition (c). Example 5.1. Let a function f : U --+ IR n satisfy the hypotheses of Theorem 5.1. In this theorem we have shown that the space DU) satisfies condition (e). Theorem 5.1 and Lemma 5.1 imply that it satisfies condition (u) too. When we apply Lemma 3.4 and Theorem 4.2 we obtain the closedness of the space DU) in the solution space of the inequality
CHAPTER 6
204
Ily'(t)II :::; CPo(t). The Corollary of Theorem 4.1, Theorem 4.2, and Lemma 4.4 imply the fulfilment of condition (c). Thus D(J) E Rceu(U). In many situations the following assertion is helpful. Lemma 5.4. Let a single valued function h : V --t ~n be defined and be continuous on an open convex subset V of the space ~I and have continuous partial derivatives (:~ are vector functions with values in the space ~n). Let the derivatives be bounded in norm by a number M. Then for every two points p = (PI, ... ,PI) and q = (q1, ... , ql) of the set V we have Ilh(p) - h(q)11 :::; Milip - qll (i.e., the function h satisfies the Lipschitz condition with the constant MI). Proof. Let
cp(s)
=
d ds (h(q
+ s(p -
q»)
(
I 8h =~ 8Xi (q + s(p -
) q»(Pi - qi) .
We have
h(p) - h(q)
=
J1
cp(s)ds
o
I (1J::i (q + s(p - q»ds) (Pi - qi)·
=~ 0-1
0
Since II :~ II :::; M, Ilh(p) - h(q)11 :::; Milip - qll. The lemma is proved. • Example 5.2. Let a single valued continuous function f : U --t ~n have continuous partial derivatives in y. Show that D(J) E Rceu(U). Take an arbitrary point (to, Yo) E U. Find c > 0 such that [to -c, to +c] x [Goyo] ~ U. Let M = sup{llf(x)II, II -tt-(x) II, ... , II /t(x)11 : x E [to - c, to + c] x [GoYol} « 00, see Remark 2.3.4), V = (to - c, to + c) X Geyo. The function flv satisfies condition a) and the first part of condition b) (with CPo == M). By Lemma 5.4 it satisfies the second part of b) (with CP1 == Mn) too. Theorem 5.1 may be applied to the equation y' = f(t, y) in the region V. Therefore D(J) E Reu(U) (see also Lemmas 5.1,5.2 and 5.3). By reasoning of Example 4.1 D(J) E Rc(U). Thus D(J) E Rceu(U). It is not difficult to mention concrete equations satisfying the restriction of Examples 5.1 and 5.2. These examples give a sufficient number of situations where our near considerations are valuable.
6. Maximal extensions of solutions Let U be an open subset of the product ~ x ~n. For Z E R(U) denote by Z+ (respectively, by Z-) the set of all continuous functions z : [a, b) --t IRn (respectively, z : (a, b] --t IRn , 00 and -00 may be open endpoints) satisfying the conditions: zlI E Z for every segment I ~ 7r(z) and there is no element of the space Z extending the function z.
Basic properties of solution spaces.
205
Denote by Z-+ the set of all continuous function Z : (a, b) ---t ]Rn such that for some (then by virtue of the condition Z E R{U) for every) t E (a, b) we have zl(a,t] E Z- and Zl[t,b) E Z+. Lemma 6.1. Let Z E Rc(U). Let a function z : [a, b) ---t ]Rn belong to Z+. Then the graph of the function z goes out from every compact subset of the set U as t ---t b, i.e., for every compactum K ~ U there exists a number to E [a, b) such that (t, z(t» E U \ K for t E (to, b). Proof. I. Notice that there is no compact subset A of the set U containing Gr(z). In the opposite case the equality b = 00 is impossible by virtue of the compactness of the set A. Next, by virtue of the compactness of the set ZA the sequence {zl[a,b-2- k (b_a)]: k = 1,2, ... } has a subsequence converging to a function Zo E Z A. The function Zo extends the function z. This contradicts the condition z E Z+. II. Let us go on to the main part of the proof. Let K be an arbitrary compact subset of the set U. Assume that the compactum K does not satisfy the mentioned condition. By Lemma 2.4.2 there exists a compact subset A of the set U containing in its interior the compactum K. By I we have a possibility to choose points S1 < t1 < S2 < t2 < S3 < t3 < ... of the half interval [a, b) in a manner that (Si' Z(Si» E K, (ti' z(t i E vA and Gr{zi) ~ A, where Zi = ZI[Si,ti)" By virtue of the compactness of the set ZA ;> {Zi: i = 1,2, ... } the sequence {Zi: i = 1,2, ... } has a subsequence converging to a function Z· E Z A. By virtue of the continuity of the mapping 7r, 7r(z*) consists of one point limi-+oo Si = limi-+oo t i . On the other hand its graph must meet both the sets K and vA. The sets are disjoint. So our assumption is false. The lemma is proved. • Evidently analogous assertions are true for Z- and Z-+ too. Lemma 6.2. Let Z E Rce(U) and Zo E Z. Then there exists a function Z E Z-+ extending Zo. Proof. Let )..(t,y) = {7r{z): Z E Z, SUP7r(z) = t, z{t) = y} for (t,y) E U. The set )..(t,y) consists of segments with the right endpoints t. The union M = U)..(t, y) is connected (see Example 2.6.1 and Theorem 2.6.1). By Theorem 2.6.3 the set M is either a segment, either a half interval, or an interval. The set M contains the point t and does not meet the interval (t,oo). Thus M is either a segment or a half interval with the right endpoint t E M. The space Z satisfies condition (e). This implies immediately the openness of the set M in (-00, t]. Thus the set M is a half interval (a{t, y), t]. Construct now successive extensions Zi, i = 1,2, ... , of the function Zo in the left. Assume that we have a function Zi, i = 0,1,2, .... Let [ai, bo] = 7r{Zi)' The definition of a{t, y) implies the existence of a function z;+1' 7r{z;+1) = [ai+1,ai], taking the value zi{ai) at ai; moreover, the left endpoint ai+1 of 7r{z;+1) belongs to (a{ai' zi{ai», a(ai' zi(ai) + 2- i ) if
»
206
CHAPTER 6
a(ai' Zi(ai)) > -00, and satisfies the inequality ai+! < -i - 1 if we have a(ai' zi(ai)) = -00. Define the function for t E [ai+!, ail, for t E [ai, bo], 7r(Zi+!) = [ai+!, bolo Evidently under this construction
and if a = limi->oo a(ai' zi(ai)), then a = limi->oo ai' The functions Zi, i = 1,2, ... , sequentially extend each other. We have the possibility of defining a function z* on the half interval (a, bol = U~l[ai,bol by putting z*(t) = Zi(t) for t E [ai,bol. Decuase of our construction the restriction of the function z* to every segment I ~ (a, bol belongs to Z (because I ~ [ai, bol for some i = 1,2, ... ). Assume that there exists a function Z E Z extending the function z*. Since Z E Re(U), we can assume in addition that a E (7r(z)). So the points a(ai' zi(ai)) and ai lie in (7r(z)) beginning with some i:
This contradicts the definition of the number a(ai' zi(ai))' So z* E Z-. Likewise we extend the function Zo in the right to z** E Z+. The function: Z* (t) for t E 7r(z*), z(t) = { z**(t) for t E 7r(z**), is defined on the interval7r(z*)U7r(z**) and satisfies all imposed conditions. The lemma is proved. • Lemma 6.3. Let Z E Rc(U). Then for every compactum K ~ U there exists a number 8 > 0 such that for every function Z E ZK and every extension Zo E Z-+ of the function Z we have Oo7r(z) ~ 7r(zo) (i.e., if 7r(z) = [a, bl, then (a - 8, b + 8[~ 7r(zo)). Proof. By Lemma 2.4.2 there exists a compactum A ~ U containing the compactum K in its interior. Assume that the stated condition is not fulfilled. Lemma 6.1 implies that for every k = 1,2, ... there exists a function Zk E Z A with the graph meeting both the sets K and 8A and with l7r (Zk) < 2- k • By virtue of the compactness of the set Z A the sequence {Zk: k = 1,2, ... } contains a subsequence converging to a function z* E ZA' The continuity of the mapping 17r implies that 17r(z*) = O. Hence 7r(z*) consists of one point only. On the other hand, the graph of the function z* must meet both the sets K and 8A (by virtue of our definition
Basic properties of solution spaces.
207
of the topology of Cs(U) the sets {z: z E Cs(U), Gr(z) n K =F 0} and {z: z E Cs(U), Gr(z) n 8A =F 0} are closed). Our assumption has led to a contradiction. The lemma is proved. • 7. Continuity of the dependence of solutions on initial values As before let U denote an open subset of the product lR x lRn. Theorem 7.1. Let Z E Rc(U). Let K be a compact subset of the set U. Let G(x) = {z: z E ZK, Gr(z) 3 x} for x E K. Then the (multi-valued) mapping G : K --t ZK is upper semicontinuous. Proof. The product K x ZK is compact (Theorem 2.1.3). Theorem 3.3.2 implies the closedness of the set A = {(x,z): x E K, z E ZK, Gr(z) 3 x} in the product K x ZK (see also Theorem 1.5.3). Theorem 1.6.1 implies its compactness. Apply Corollary of Theorem 1.7.8 to the restriction to the compactum A of the projection of the product K x ZK into the first factor K. It implies that the multi-valued mapping g associating to a point x E K the set ({x} x Z K) n A is upper semicontinuous. The mapping G is the composition of the mapping g and of the projection of the product K x ZK onto the second factor Z K. This implies the upper semicontinuity of the mapping G. The theorem is proved. • Theorem 7.2. Let hypotheses of Theorem 7.1 hold. Let I be a segment of the real line. Let G 1 (x) = G(x) n 7["-1(1) for x E K. Then the (multivalued) mapping G 1 : K --t ZK is upper semicontinuous. Proof. The proof reduces to referring to Theorems 7.1 and 2.5.3. (We apply the last one to the mapping G and the constant mapping F == Z K n
7["-1(1)).
•
Corollary. Let with the notation of Theorem 7.2 for every point x of a set X ~ K the set G 1 (x) consist strictly of one function. Then the (single • valued) mapping G1 1x is continuous. In this corollary we obtain the continuity of the dependence of solutions of the Cauchy problem on initial values in the sense which is close to the most common understanding of these words. The closeness turns into full coincidence when in the definition of the mapping G 1 we replace the compactum K to the entire set U, for instance, if it is known in advance that the graphs of all solutions of the Cauchy problem with given and near initial values lie in a compact urn K ~ U j and so the mention of the compactum K in the definition of G 1 does not carry any real information. Theorem 7.3. Let Z E Rce(U) and (to, Yo) E U. Then there are numbers J.L > 1/ > 0 such that the set K = [to - 1/, to + 1/J x [OJLYoJ lies in U and for every function z E Z-+, the graph of which meets the set A = [to - 1/, to + 1/J x [OIlYO], the conditions [to - 1/, to + 1/J ~ 7["{z) and z([to - 1/, to + 1/]) ~ OJLYo hold.
208
CHAPTER 6
Proof. Take a number 1£ > 0 such that the compactum
P = [to - 1£, to
+ 1£]
x [0I'Yo]
lies in U. Find a number 8> 0 according to Lemma 6.3 for the compactum P. By Theorem 3.6.1 the set Zp is equicontinuous. So there exists a number v E (O,minH, ~}) such that IIZ(8) - z(t)11 < ~ for every function z E Zp and every points 8, t E 7r(z) satisfying the condition 18 - tl < 2v. Show that the numbers 1£ and v satisfy the imposed conditions. Let the graph of a function Z E Z-+ meet the set A and for some t E [to - v, to + v] n 7r(z) the point z(t) belong to the set [O"Yo]. By our choice of the numbers 8 and v we have [to - v, to + v] ~ (t - 8, t + 8) ~ 7r(z). Let J(8) denote the segment with the endpoints 8 and t. By the choice of
v, if 8 E [to - v, to
+ v]
and Gr (zIJ(s))
Ilyo - z(8')11 ~ lIyo - z(t)11
~ P, then
+ Ilz(t) - z(8')11 < ~ + ~ = 1£
for every 8' E J(8). Therefore Z(J(8)) ~ 0I'Yo. Now the condition Z E Re(U) implies the openness of the set
M = {8: 8 E [to - v, to
+ v],
Gr(zIJ(s)) ~ K}
in the segment [to - v, to + v]. By virtue of the continuity of the function z the set M is closed in the segment [to - v, to + v]. The connectedness of the segment implies that M = [to - v, to + v]. Therefore z([to - v, to + v]) ~ 0I'Yo. The theorem is proved. • Let us return to the discussion of Theorem 7.2 and its Corollary. In Theorem 7.3 for an arbitrary point (to, Yo) of the set U we find its neighborhood (A), a compactum K and a segment I = [to - v, to + v] such that the graph of every solution of the Cauchy problem on the segment I with initial values from (A) lies in K. Therefore the mentioning of the compactum K in the definition of G 1 may be omitted. Thus in this case our assertions establish, respectively, the upper semicontinuity and the continuity of the dependence on initial values of solutions of the Cauchy problem on the segment I with initial values from (A). If in addition Z E Ru(U) then we can take the set A as the set X in Corollary of Theorem 7.2. This relates to, in particular, the solution spaces of equations in Examples 5.1 and 5.2.
CHAPTER 7
CONVERGENT SEQUENCES OF SOLUTION SPACES
In this chapter we discuss the continuity of the dependence of solutions of an equation on parameters of the right hand side. The notion mentioned in the title of the chapter is a convenient tool of investigation. Further we will see that we can apply it not only in the discussion of the continuity but in the investigation of other properties of solution spaces, for instance, in the proof of the existence theorems. In particular, remarks of §2 below, which are not used in practice in the classical theory of Ordinary Differential Equations, are strong tools in the investigation of properties of solutions. We can prove the convergence of sequences of solution spaces in framework of the classical theory with the help of the corresponding versions of theorems on the continuous dependence of solutions on parameters (see, for instance, Theorem 1.2.4 in [Hal). In this chapter we will show how we can prove the inessentiality of singularities of the right hand side and keep intact the continuity of the dependence of solutions on the parameters. We introduce here the topological space Rc(U). It is closely related to the concept of convergent sequences of solution spaces. Both the notion mentioned in the title of the chapter and the topological space Rc(U) are the main topological notions of our theory. 1. Continuity of the dependence of solutions of equations on parameters of right hand sides
In idea, the property which we will speak about, looks as follows. We have a differential equation y' = f(t, y, J-L). The right hand side of the equation depends on a parameter J-L. We also have a point (to, Yo) E U and a segment I 3 to. For every value of the parameter J-L let the Cauchy problem y' = f(t, y, J-L), y(to) = Yo on the segment I have an unique solution z/-" Then the mapping associating the function z,.. to the parameter J-L must be Continuous. However, here we immediately meet the question about the existence and the uniqueness of solutions of the mentioned Cauchy problems. First, such a solution need not exist even under the hypotheses of Theorem 6.5.1
CHAPTER 7
210
if the segment I is too large. Secondly, if a solution exists, in contrast to the situation of Theorem 6.5.1 it need not be unique. Therefore the in general case we can speak about any analog of the continuity of the corresponding multi-valued mapping. The consideration of simplest examples shows that the upper semicontinuity suits here. The classical theory suppose that the right hand side of the equation y' = f(t, y, /L) is continuous in the totality of arguments (Le., with respect to the Tychonoff topology of the product U x M, where U is a open subset of the space ~ x ~n, M is the set of values of the parameter /L). When we assume that the set of parameters M is a metric space the condition of the continuity of the function f(t, y, /L) in the totality of arguments turns into the condition: (1.1) for every compactum K ~ U and for every sequence of parameters /Lj ~ /Lo the sequence of functions CPj(t,y) = f(t,y,/Lj), (t,y) E K, converges uniformly to the function CPo (defined by the same formula). (By virtue of Theorems 3.2.1 and 3.2.4, Coroll<;try of Lemma 1.7.1 and Corollary of Theorem 3.5.4 these conditions are equipotent.) When we discuss here technical aspects of the continuity of the dependence of solutions on parameters in connection with (1.1), we need not consider the entire set M at once. We can consider a particular sequence of parameters /Lj ~ /Lo and in statement of results we need not mention its members explicitly, Le., we can consider a sequence of functions CPj converging to a function CPo in the sense of a corresponding version of condition (1.1), namely with respect to the compact open topology on the space C(U, ~n) (see Theorem 3.2.3). Further applications suggest that we should not restrict our investigation by the above situation. We will obtain a little stronger result, although with a complication in the proof. This complication is related not to the complication of constructions in this section. It is related to the use of the larger material amount in the previous chapters. See also [Ar]. Let F o, F : U ~ ~n be multi-valued mappings, K be a compact subset of the set U, z E Z ~ Cs(U). Put
oz(Fo, F) = inf
{rl1r(cp) cp(t)dt:
7r(cp) = 7r(z) and F(t, z(t))
cP is a positive measurable function,
~ Ocp(t)Fo(t, z(t)) for almost all t E 7r(cp) }
and oz,K(Fo,F) = sup{oAFo,F): Z E ZK}. We say that a sequence of multi-valued mappings Fi : U ~ IR n , i = 1,2, ... , converges to a mapping Fo : U ~ IRn with respect to a
Convergent sequences of solution spaces.
211
space Z ~ Cs(U) (in the set U) if oz,K(Fo, F i ) --+ 0 for every compact subset K of the set U. We say that a multi-valued mapping F : U --+ jRn is weakly continuous with respect to a space Z ~ Cs(U) if for every segment la, bj, -00 < a < b < 00, of the real line and for every sequence {Zk: k = 1,2, ... } ~ Z, 7r(Zk) = la, b], converging uniformly to a function Z E Z:
if Denjoy integrable functions o.k, k = 1,2, ... are defined on a segment la, bj, o.k(t) E F(t, Zk(t)) for almost all tEla, bj and the sequence of the functions
{A,(t) =
i
a,(s)ds: k=1,2, ... }
converges uniformly to a function A then the function A is generalized absolutely continuous and A'(t) E F(t, z(t)) for almo.st all tEla, bj.
Example 1.1. Let f : U --+ jRn be a single valued continuous function. Then the function f is weakly continuous with respect to the space Cs(U): by virtue of the single valuedness of the function f every function o.k, which occurs in the definition, coincides on a set of full measure with the continuous function rpk (t) = f(t, Zk (t)), and the fulfilment of the required condition follows from the continuity of the integral with respect to the metric the uniform convergence (see, for instance, Theorems 3.3.1 and 3.7.2). If a mapping F : U --+ jRn is weakly continuous with respect to a space Z and Zl ~ Z then the mapping F is weakly continuous with respect to the space Zl' This follows immediately from our definition. Denote by Q/(U) the set of all mappings F E Q*(U) with nonempty compact values satisfying the condition if K ~ jRn is a compact, I is a segment of the real line, and I x K then the mapping Fl 1xK satisfies condition (4.13.2).
~
U
The graph of every function Z E Cs(U) (by virtue of its compactness) may be covered by a finite number of sets of the type mentioned in this condition. Thus Corollary of Theorem 4.13.1 and Theorem 4.12.2 imply: Lemma 1.1. Let F E Q/(U) and Z E Cs(U). Then the mapping cp: 7r(z) --+ jRn, rp(t) = F(t,z(t)), is measurable. • Lemma 1.2. Let F : la, bj --+ jRn be a measurable multi-valued mapping of a segment la, bj, a < b, of the real line into the space jRn. Let values of the mapping F be closed. Let 'l/J : la, bj --+ jRn be a measurable single valued /unction, rp : la, bj --+ jR be a measurable single valued n~m-negative function and G(t) = F(t) n O/('l/J(t), rp(t)) for t E [a, b]. Then the mapping G is measurable.
212
CHAPTER 7
Proof. The proof consists in referring to Lemmas 4.12.3 and 4.12.4 . • Lemma 1.3. Let F E Q/(U), Z E C.(U). Let
such that the sequence {Zi: i = 1,2, ... } converges to the function z. Fix positive integrable functions
J
~ O"'i(t)F(t, Zi(t)).
71"( z;)
By Lemma 1.3 there exist measurable functions f3i : 71'(Zi) - JRn satisfying the conditions f3i(t) E F(t, Zi(t)) and IIz;(t) - f3i(t) II ~
Convergent sequences of solution spaces.
213
The right hand side of the inequality
IIBi(t) - (Z(t) - z(a + c:))11 ~
~
IIBi(t) - (Zi(t) - Zi(a
+ c:))11 + II(Zi(t)
- Zi(a + c:)) - (Z(t) - z(a + c:))11
t
J
I"Si(S) - z:(s)lIds
+ IIZi(t)
- Z(t) II
+ IIZi(a + c:) -
z(a + c:))11
tends to zero as i - t 00 (for a + c: ~ t ~ b - c:). This estimate does not depend on t. We have IIBi - (z - z(a + c:))11 - t 0, i.e., Bi - t Z - z(a + c:). By virtue of the weak continuity of the mapping F with respect to the space Z the inclusion (z(t) - z(a + c:))' E F(t, z(t)) is true for almost all t E [a + c: < b - c:l, i.e., z'(t) E F(t, z(t)). Hence ZI[a+e,b-e] E D(F). Now the arbitrariness in the choice of c: implies that Z E D(F). The theorem is proved. • The result obtained does not give a possibility to analyze completely the situation described in the beginning of the section. In order to make a last step introduce the set s(U) of all sequences {Zk: k = 1,2, ... } ~ Ri(U) satisfying the condition: for every compact subset K of the set U and for every infinite set A ~ N every sequence Zk E (Zk)K' k E A, is equicontinuous. We say that a sequence {Zk: k = 1,2, ... } of subspaces of the space C8 (U) converges to a space Z ~ C.(U) in U if for every compact subset K of the set U and for every sequence of functions Zk E (Zk)K' k E A, there are a function Z E Z and a subsequence {Zk: k E B} of the sequence {Zk: k E A} converging to the function z: Zk - t Z as k E B, k - t 00 (notice that when we introduce this definition we prepare the using of Theorem 2.5.1). In what follows we will allow the use of these notions and notation in more general situations, when U is a locally compact subset of the product R X X of the real line lR and of a metric space X. In particular, Remark 1.1 and Lemma 1.4 below concern just this general case. Theorem 3.6.1 implies easily: Remark 1.1. Every convergent sequence {Zk: k = 1,2, ... } ~ Ri(U)
belongs to s(U). On the other hand: Lemma 1.4. Let a
= {Zk: k = 1,2, ... } E s(U) and lim top sup Zk lie in Z ~ C 8 (U). Then the sequence a converges in U to Z. k-+oo
214
CHAPTER 7
Proof. Let K be an arbitrary compact subset of the set U. Consider a sequence of functions Zk E (Zk)K' k E A. By virtue of the condition {Zk : k = 1,2, ... } E s(U) and Theorem 3.7.1 we can pass to a subsequence {Zk: k E Ad converging to a function Z E lim topsUPk--+oo Zk. This gives what was required. The lemma is proved. • Lemma 1.4 and Theorem 1.1 imply immediately: Theorem 1.2. Let hypotheses of Theorem 1.1 hold and {Z n D(Fk ) : k = 1,2, ... } E s(U). Then the sequence {Z n D(Fd: k = 1,2, ... } converges in U to the space Z n D(F). • When we compare the last result with Theorem 2.5.1 we obtain the possibility of proving the upper semicontinuity of the mapping which associates to a mapping F : U ~ IR n the set D(F, K)j see §6.2. The restriction that we consider solutions with graphs lying in a compactum K is not very essential. By this we mean that if we restrict the consideration to solutions with small domains of definition then, as rule, we can point a compact subset of the set U which contains the graphs of all solutions of the Cauchy problem with given and near initial values (see Theorem 6.7.3 and Lemma 8.1.1). Let us return to the situation discussed in the beginning of this section. Theorem 1.3. For every compact subset K of the set U the multi-valued mapping q,K: C(U,lRn) ~ Cs(U), q,K(f) = D(f,K), is upper semicontinuous. Proof. By virtue of the metrizability of the spaces C(U, IRn) (Theorem 3.2.2) and Cs(U) we can use Theorem 2.5.1. Consider an arbitrary sequence of functions {h: j = 1, 2, ... } ~ C (U, IRn) converging in the space C (U, IRn) to a function fo. Let Zj E q,K(fj) for j = 1,2, .... By Lemma 2.4.2 there exists a neighborhood V of the compactum K, the closure of which is compact and lies in U. By Theorem 3.2.3 we have the uniform convergence fjl[V] ~ fol[V]. By Theorems 3.5.3 and 1.7.8 (see the proof of Theorem 3.6.1) and by the definition of the topology of the space of partial mappings we obtain the compactness of the set U{Gr(fjl[V]) : j = 0,1,2, ... }. Lemma 2.3.6 implies the existence of a number m ~ 0 such that IIfj(x)II ~ m for every j = 0,1,2, ... and x E [V]. Thus the set U{D(fj, V): j = 0,1,2 ... } lies in the solution space of the inequality Ily'(t)II ~ m. By Lemma 6.3.1 {D(fj, V): j = 0,1,2, ... } E s(V). The uniform convergence fjl[V] ~ fol[V] implies the convergence with respect to the space Cs(V). In view of the remarks of Example 1.1 referring to Theorem 1.2 completes the proof. The theorem is proved. • In Theorem 1.3 we have proved the continuity of the dependence of solutions of equations on the right hand sides (with respect to the compact open topology of the space C (U, IRn)). Although here we do not mention any parameters of the right hand side, we immediately obtain a corresponding
215
Convergent sequences of solution spaces.
wording 'with parameters' when we compare Theorem 1.3 with Theorems 3.2.3 and 3.2.4. Theorem 1.3 may be added by an assertion about the continuity of the dependence of solutions on initial values. To do it we prove first the following auxiliary assertion. Lemma 1.5. Let F : X -+ Cs(U) be a multi-valued mapping of a metric space X into the space C s (U). Let for every compact subset K of the set U the mapping FK : X -+ Cs(U), FK(X) = F(x) n Cs(K), have compact values and be upper semicontinuous. Then the multi-valued mapping
G : X x expc U x expc U -+ Cs(U), G(x, K 1 ,K2) = {z: z E F(x), Gr(z) n Kl -=1= 0, Gr(z)
~
K 2},
is upper semicontinuous. Proof. By virtue of the metrizability of the domain of the mapping G and of the space Cs(U) we can use Theorem 2.5.1. Consider an arbitrary sequence {(Xj, Kf, K4): j = 1,2, ... } of points of the domain of the mapping G converging to a point (x, K 1 , K2). Let Zj E G(xh Kf, K4) for j = 1,2, .... By Lemma 2.1.2 Xj -+ x, Kf -+ Kl and K4 -+ K 2. By Lemma 2.4.2 there exists a neighborhood V of the compactum K 2 , the closure of which is compact and lies in U. By virtue of the convergence K4 -+ K2 we have K4 ~ V, beginning with some j = jo. Hence Gr(zJ ~ V. Apply Theorem 2.5.1 to the mapping F rv ). Since Zj E Frv)(xj), the sequence {Zj : j = 1,2, ... } contains a subsequence converging to a function Z E F(V)(x). Theorem 3.3.2 implies that Gr(z) ~ K 2 . Lemma 2.5.3 implies the nonemptiness of the intersection Gr(z) n K 1 . By Theorem 2.5.1 the lemma is proved. • Theorem 1.3 and Lemma 1.5 imply immediately: Theorem 1.4. The multi-valued mapping W : C(U, IRn) x expc U x expc U
w(f, K 1 , K 2 )
= {z: z E D(f, K 2 ),
Cs(U), Gr(z) n Kl -=1= 0} -+
•
is Upper semicontinuous.
2. Properties of convergent space sequences Lemma 2.1. Let a sequence {Zk: k a space Z E R(U). Let Zk E z;+ for beginning with some k = ko, and the COnverge uniformly to the function z. oj the function Z and a subsequence
= 1,2, ... } ~ Rc(U) k = 1,2, .... Let Z E
converge in U to Z, 7r(z) ~ 7r(Zk),
sequence {zkl1r(z): k = ko, ko + 1, ... } Then there are an extension z* E Z-+ {Zk: k E A} such that I ~ 7r(Zk) for
216
CHAPTER 7
every segment I ~ 7r(z*), beginning with some k = kl E A, and the sequence {zkII: k E A, k ~ kd converges uniformly to the function Z*II. Proof. I. Show first that for some 8 > 0 there exists a subsequence {Zk : k E B} of the sequence {Zk: k = 1,2, ... } such that 7r(Zk) ;2 [a - 8, b + 8], where k E B, [a, b] = 7r(z), and the subsequence converges uniformly on [a - 8, b + 8] to an extension Zo E Z of the function z. Fix a compactum K ~ U which contains the set Gr(z) in its interior. By virtue of the convergence Zk I... (z) ---t z, by Theorem 3.5.3 and by the definition of the topology of the space Cs(U) we have Gr(zkl ... (z») ~ (K), beginning with some k = ko. Lemma 6.6.1 implies the existence of points ak < bk of the interval7r(zk) such that [a,b] ~ [ak,b k], Gr(zkl[ak.bkl) ~ K and (ak' zk(ak)), (b k , zk(bk)) E 8K. Because of the convergence of the sequence {Zk: k = 1,2, ... } to Z in U the sequence {Zk = Zkl[ak,bkl: k = ko, ko + 1, ... } has a subsequence {zk: k E A *} converging to a function Zo E Z K. By Theorem 3.5.1 and its Corollary 7r(Zk) ---t 7r(zo) or ak ---t ao and bk ---t bo, where lao, bo] = 7r(Zo), k E A*. By Theorem 3.5.4 (ak' zk(ak)) ---t (ao, zo(ao)) and (bk, zk(b k )) ---t (b o, zo(b o)) (k E A*). Hence (ao, zo(ao)), (b o, zo(b o)) E 8K. The function Zo is the limit of a sequence converging on the segment 7r(z) to the function z. Thus zo extends z. Therefore the graph of Zo on the segment 7r(z) ~ n{[ak' bk] : k E A *} lies in (K). Hence the points ao and bo do not belong to the segment 7r(z). So 7r(z) ~ (ao, bo). Take as 8 an arbitrary positive number which is less than the distance from the points ao and bo to the segment 7r(z). For this choice of 8 the segment [a - 8, b+ 8] lies in the segment [ak' bk], k E A*, beginning with some k = kl E A* (use the continuity of the mapping 7r). The convergence of the sequence {Zk I[a-6,bHl: k E A*, k ~ kd to the function ZOI[a-6,bHl follows now from Theorem 3.5.4. II. Now we repeat with some modifications and complications the proof of Lemma 6.6.2. Choose subsequences:
1m
= {Zk:
k E Am}, m
= 1,2, ... ,
({I, 2, ... }
= Ao
;2
Al ;2 A2 ;2 ... )
and fix numbers 'T/o( = a) ~ 'T/l ~ 'T/2 ~ ... in such a manner that on the segment ['T/m, bJ the sequence 1m converges uniformly to a function z;:: E Z. Let Am be the set of all segments [c,'T/mJ, on which we can choose a subsequence of the sequence 1m converging to an element of the space Z. Theorem 2.6.1 implies the connectedness of the set M = UA m. Because of! the set is open in the half interval (-OO,'T/mJ. By Theorem 2.6.3 the set M is a half interval with the open left endpoint: M = (am,'T/mJ, where am < 'T/m· Take as'T/m+l an arbitrary number from (am,min{am + 2- m ,'T/m}) if am > -00, and from (-00, min{ -m, 'T/m}) if am = -00. By our definition
Convergent sequences of solution spaces.
217
of am there exists a subsequence I'm+! of the sequence I'm converging on the segment [1]m+l,1]mj to an element of the space Z. The sequence I'm converges on the segment [1]m, bj. Thus the sequence I'm+! converges on the segment [1]m+!, bj (to an extension z;';;'+! E Z of the function z;;). Under this construction
For 1] = limm-+oo am we have 1] = limm-+oo 1]m' Define the function z** on the half interval (1], bj = U{[1]m, bj m = 1,2, ... } by putting z**(t) = z;;(t) for t E [1]m, b], m = 1,2, .... Now fix an arbitrary increasing sequence of indices A' = {k m m = 1,2, ... }, where km E Am. If a segment I lies in (1], b], then I ~ [1]m, b] for some m = 1,2, .... On the segment I the subsequence {zkll: k E A', k ~ k m} of the sequence {zkll : k E Am} converges uniformly to the function Z"'II = Z;';;'II' Show that z ... E Z-. Assume the opposite. This holds only if there exists an extension z E Z of the function z**. By Lemma 2.4.2 there exists a compactum K ~ U containing in its interior the graph of the function z. Repeat the construction as in I, but for the sequence {Zk: k E A'} instead of the initial sequence 1'0' The function Zo appearing in this construction extends the function z**. The endpoints ao < bo of 7r(z**) satisfy the condition ao < 'f] < b < boo If 8* is an arbitrary positive number less than 'f] - ao and bo - b, then (ao + 8*, bo - 8*) ;2 ['f], b]. We have lao + 8*, bo - 8*] ~ [ab bk] beginning with some k = kl E A' (to prove the latter fact use the continuity of the mapping 7r as in I). Theorem 3.5.4 implies the convergence of the sequence {zkl[ao+6*,b o-O*j: k E A'} to the function ZOI[ao+6*,bo-O*j. We have 'f] = limm-+oo am E (ao + 8*, bo - 8*). So Q m E (ao + 8*, bo - 8"), beginning with some m = mo. This contradicts the definition of am, m = mo, mo + 1, mo + 2, .... Thus our assumption is false and z** E Z-. III. By analogy with II construct an extension in the right z*** E Z+ of the function z and choose in the sequence {Zk: k E A'} a subsequence {Zk: k E A} converging uniformly on every segment I ~ 7r(z***) to a function z"'*. Define on the interval 7r(z"') U 7r(z"'*) the function: z*
z**(t)
{ (t ) -- z ..... (t)
for t E 7r(z**), for t E 7r(z·"').
The verification of the fulfilment of all imposed conditions is not difficult. The lemma is proved. • The following assertion is convenient for the proof of the existence the; orems (Le., for the verification the fulfilment of condition (e) ) for solution
218
CHAPTER 7
spaces of equations with various types of right hand sides. In the next chapter we meet examples of such situations. But the reader now has Theorem 6.5.1 (see also Examples 6.5.1 and 6.5.2) and he may use this result for independent exercises. Theorem 2.1. Let a sequence {Zk: k = 1,2, ... } ~ Rce(U) converge in U to a space Z E R(U). Then Z E Re(U). Proof. Take an arbitrary point (t, y) of the set U. Denote by Z the function defined on the one point set {t} and taking the value y at t. The membership Zk E Re(U) implies that Z E Zk for every k = 1,2, .... By virtue of the convergence of the sequence {Zk: k = 1,2, ... } in U to the space Z the latter fact implies the belonging of the function Z to the space Z. By Lemma 6.6.2 for every k = 1,2, ... there exists an extension Zk E Z;;+ of the function z. When we apply Lemma 2.1 we obtain the existence of a function z* E Z-+, the domain of which contains the point t and which takes the value y at t. This implies immediately what was • required. The theorem is proved. Let us now state an assertion closely related to the continuity of the dependence of solutions on (parameters of) the right hand side and initial values. Theorem 2.2. Let a sequence {Zk : k = 1,2, ... } ~ Rce(U) converge in U to the space Z E R(U), Zk E Z;;+, tk E 7r(Zk). Let the sequence {(tk,Zk(t k )): k = 1,2, ... } converge to a point (t,y) E U. Then there are a function z* E Z-+, 7r(z*) 3 t, z(t) = y, and a subsequence {Zk: k E A} such that for every segment I ~ 7r(z*) we have I ~ 7r(zd beginning with some k = ko E A, and the sequence {Zk II: k E A, k ~ k o} converges uniformly to the function z*II. Proof. Fix a compact subset K of the set U containing the point (t, y) in its interior. Then (tk' Zk(t k )) E (K) beginning with some k = k 1 . Lemma 6.6.1 implies that there are numbers ak < bk such that tk E [ak' bk ] ~ 7r(Zk), Gr (zkl[ak,h]) ~ K and (ak, zk(ad), (b k , zk(bd) E oK. Show that t E [ak' bk] beginning with some k = k2 > k 1 . Assume the opposite. The convergence of the sequence {Zk: k = 1,2, ... } to the space Z in U and our additional assumption imply that we can choose a subsequence {zkl[akobk] : k E Ao} converging to a function Zo E ZK. We have t tt [ak' bk] for every k E Ao. Since It - tkl ---+ 0, we can require in addition the fulfilment of one of the conditions It - ak I ---+ or It - bkI ---+ 0, k E Ao· Theorem 3.5.4 implies that (t, zo(t)) = (t, y) E oK. This contradicts the choice of the compactum K. The obtained contradiction gives what was required. The sequence {zk = zkl[ak,bk]: k = k2' k2 + I, ... } contains a convergent subsequence {Zk: k E Ad. Theorem 3.5.4 implies that Zk(t) = zZ(t) ---+ y, k E A l . Now the application of Lemma 2.1 to the sequence {Zk: k E Ad
°
Convergent sequences of solution spaces.
219
and to the function z = zol{t} gives what was required. The theorem is proved. • Theorem 6.7.3 is convenient for the comparison of local and global properties of solution spaces of ordinary differential equations. In this context it is helpful to have the following assertion. Theorem 2.3. Let Z E R(U). Let V be an open subset of the set U, Zv E Rce (V). Let K be a compact subset of the space IRn. Let OK be its neighborhood (in the space IRn), I be a segment of the real line, and I x OK ~ V. Assume that if the graph of a function z E (Zv)-+ meets the set I x K, then I ~ 7r(z) and z(I) ~ OK. Then if the graph of a function Zo E Z-+ meets the set I x K, then I ~ 7r(zo), zolr E Zv and zo(I) ~ OK. Proof. Let Zo E Z-+, tl E In 7r(zo), zo(it) E K. Denote by M the set of all s E I such that the segment J(s) with the endpoints it and s lies in 7r(zo) and z(J(s» ~ OK. Evidently the set M is either a segment either a half interval or an interval. The point tl lies in (M) [" Let a < b be the endpoints of M. Fix numbers
in a manner that a = limk--+oo ak and b = limk--+oo bk . By Lemma 6.6.2 for every k = 1,2, ... there exists an extention Zk E (Zv)-+ of the function ZOhak,bkj' Apply Lemma 2.1 with Zk == Zv for k = 1,2, ... , z = ZOI{td' Lemma 2.1 implies the existence of an extention z* E (Zv)-+ of the function ZOIM' Then I ~ 7r(z*), z*(I) ~ OK and the compactum Gr(z*lr) (~ V) contains Gr (ZOIM)' By virtue of the membership Zv E Rc(V) this implies the closedness of the set M (in the segment I). Since
zo(M) = z*(M)
~
z*(I)
~
OK
the set M is open in the segment I. By virtue of the connectedness of the segment (see Example 2.6.1) we have M = I. This gives what was required. • The theorem is proved.
3. Equicontinuity condition Let U denote (as before) an open subset of the product IR x IRn. We saw
in §1 that the proof of the continuity of the dependence of solutions of an equation on parameters reduces to the proof of the convergence of seqUences of solution spaces. Remarks in §1 do not exhaust our possibilities in this connection. Lemma 1.4 splits the question about the convergence of a sequence a = {Zk: k = 1,2, ... } ~ R(U) to a space Z ~ Cs(U) into two: the question about the membership a E s(U) and the question about the inclusion lim topsUPk--+oo Zk ~ Z. In the next sections we point out new
CHAPTER 7
220
methods of the investigation of these questions. They differ very much from the methods of §1. Here we consider the first question. Our general plan consists in estimating the geometry of the set of singularities: on a part W of the region U the corresponding convergence (or any other property of solution spaces) it is assumed proved (for instance, with the help of versions of methods of the classical theory) and we seek in what case, and for what geometry of the set U \ W, the corresponding conclusion remains true for the entire set U. Let a multi-valued mapping (J : U --t lRn be fixed. For (t, y) E U put r~ = {{y} U H : H is a open subset of the set Ut and for some E > 0 (G~y n (J(t,y)) \ {y} ~ H} (Figure 7.1). For t E lR denote by T((J, t) the topology on the set Ut generated by the directed sets r~, y E Ut (see details in §1.2 and see the notation in II· )1 §6.4). Theorem 3.1. Let H a = {Zk: k = 1,2, ... } ~ Ri(U). For every point Figure 7.1 let (t,y) E U lim topsup{Gr (ZI[Sl,S2]) : Z E Zk, o,£-+O,k---+oo
[SI' S2] ~ (t - E, t
+ E) n 7r(Z),
Z([SI' S2]) n Gey
i= 0}
~ (J(t, y).
Let I be a family of open subsets of U. Let {(Zk)V : k = 1,2, ... } E s(V) for every V E I' For every point let t E lR the set (U \ (U'))t contain no nontrivial connected subsets on which the topology T( (J, t) and the Euclidean topology of lRn coincide (i.e., induce the same topology). Then a E s(U). Proof. Assume the opposite, i.e., that the condition a E s(U) is not fulfilled. Then there exist a compactum K ~ U and a sequence Zi E (Zi)K' i E A, which is not equicontinuous. The latter fact implies the existence of a number c > 0, of a subsequence {Zi: i E AI} and of points Si < ti from 7r(Zi), i E Al such that ti - Si --t 0 and Ilzi(S;) - Zi(t i ) II ;;:: c. The set expK is compact. Therefore the sequence {Gr(zil[Si,ti]) : i E Ad has a subsequence converging to a set F E exp K. Since
Convergent sequences of solution spaces. ti - Si
- t 0, the projection of the set F in consists of one point
t
= i-+oo,iEA lim
ti 2
~
= i-+oo,iEA lim
(in the product
221 ~
x
~n)
S i. 2
For every point y E F we have F ~ e(t,y). Therefore on the set F t the topology T(e, t) and the Euclidean topology of ~n coincide. Theorem 3.3.3 implies the connectedness of the set F. The condition IIZi(Si) - Zi(ti)11 ~ c implies the estimate diamF ~ c. Hence the set F contains more than one point. Therefore F n (U,,) =1= 0 and F n V =1= 0 for some V E 'YTake an arbitrary point y E F n V and put
8=
~min{p((t,y),(~ x ~n) \ V),c}
(see Lemma 2.3.8). Because of the convergence of the sequence {Gr(zi![Si,ti]) : i E A 2 } to the set F and by Theorem 3.4.2 we can choose a subsequence {Gr(zil[Si,t;j): i E A 3 } and points s~ E lSi, til, i E A 3 , in such a manner that (s~, Zi(S~)) E 06(t, Yl) and Zi(S~) - t y. Here Gr(zil[Si,ti]) \ 06(t, y) =1= 0, because in the case where Zi([Si, til) ~ 06Y we have:
which contradicts our choice of the functions Zi, i E A 1 • We can select points t~ E lSi, ti], i E A 3 , in such a manner that on the segment li with the endpoints s~ and t~ the graph of the function Zi lies in the compactum Kl = [06(t, y)] and (t~, Zi(tD) E {)06(t, y). The sequence {Zi!L. : i E A 3 } has no convergent subsequences, because in the opposite case the domain of the limit function would consist of one point t only and its graph would contain the point (t, y) and meet {)06(t, y) simultaneously, which is impossible. Theorem 3.7.1 implies that the sequence {zilli : i E A 3 } is not equicontinUOUs. The graphs of all its members lie in the compactum Kl ~ V. This contradicts the condition
{(Zk)V: k = 1,2, ... } E s(V). Thus our assumption is false and the theorem is proved. • Notice a particular case. The mapping e == ~n satisfies all conditions imposed above on the mapping e. The proved assertion implies: Corollary 1. Let O!
= {Zk:
k
= 1,2, ... } ~ R(U),
CHAPTER 7
222
, be a family of open subsets of the set U, for every element V of which {(Zk)V: k = 1,2, ... } E s(V). Let for every point t E IR the set (U \ (U'))t do not contain connected • nontrivial subsets. Then a E s(U). By Assertion 2.4.1 we obtain as a particular case of Corollary 1: Corollary 2. Let a
= {Zk:
k
= 1,2, ... } ~ R(U),
, be a family of open subset of the set U, for every element V of which {(Zk)V: k = 1,2, ... } E s(V). Let the set (U \ (U,)) be at most countable. Then a E s(U). • So an one point singularity does not mar the fulfilment of the condition a E s(U). In the general case the idea of Theorem 3.1 may also be realized in the following different way. Assertion 3.1. Let X and M be metric spaces, K be a compact subset of the product IR x X, cp : K - t M be a continuous mapping, and {Zi: i = 1,2, ... } ~ Ri(K), a
= {{cp(t,z(t)): z
E Zd: i
= 1,2, ... } E s(1R x cp(K)).
Then for every point (to, xo) E K the set P
z E Zj, j = k, k + 1, ... , 7r(z) ~ (to - c, to Im(z) n Ooxo # 0}]: k = 1,2, ... , c > 0, 8> O}
= n{[U{Im(z):
+ c),
lies in ({to} x X}) n cp-lcp(to, xo). Proof. Assume the opposite. Then there exists a point
By virtue of the definition of the set P this means the existence of functions Zk E Zjk and of points Sk, tk E 7r(zd, jl < h < iJ < .. , such that 7r(Zk) ~ (to - 2- k , to + 2- k ),
Z(Sk)
-t
Yo,
z(tk) - t Yl'
By virtue of the continuity of the mapping cp we have
Convergent sequences of solution spaces.
223
as k ---+ 00. Since cp(to,xo) =1= cp(to,xd, the condition a E s(lRx cp(K» is not fulfilled. The contradiction obtained gives what was required. The assertion is proved. • Assertion 3.2. Let
a = {Zk: k = 1,2, ... } ~ Ri(U),
X be a metric space, cp : U ---+ X be a continuous mapping. For every sequence Zi E Zkn kl < k2 < k3 < ... , let the sequence cp(t, Zi(t)), t E 7r(Zi), i = 1,2, ... , be equicontinuous. Let, be a family of open subsets of the set U, for every element V of which {(Zdv: k = 1,2, ... } E s(V). For every t E IR and x E X let the set (U \ (UY»t n cp-l (x) contain no nontrivial connected subsets. Then a E s(U). Proof. Assume the opposite. There are a compactum K ~ U and a sequence Zi E (Zk.)K, kl < k2 < k3 < ... , which is not equicontinuous. By
virtue of the definition of the equicontinuity we can assume in addition that 11r(zi) ---+ 0 and diamlm(zi) ~ c for some c > O. By Theorem 3.4.3 we can assume in addition the convergence of the sequence {Gr(zi): i = 1,2, ... } with respect to the Hausdorff metric to a set H E exp K. Since 11l'(zi) ---+ 0 Theorem 3.5.1 implies that the projection of the set H in the time axis (Le., in the first factor of the product IR x IR n ;2 H) consists of one point only; see Figure 7.5 below. Denote it by s. By Theorem 3.3.3 the limit H of the sequence of continua Gr(zi)' i == 1,2, ... , is connected. The estimate diamlm(zi) ~ c implies that diamH ~ c. Thus the continuum H contains more than one point. By Assertion 3.1 H ~ cp-l(X) for a point x E X. By virtue of the last hypothesis of our assertion H n V =1= 0 for some V E ,. Now we can repeat the choice of the sequence {Zi: i = 1,2, ... } and of the number c > 0 which is made in the beginning of our reasoning (see the proof of Theorem 3.1) in order to obtain the fulfilment of the following condition: there exists a compactum Kl
~
V, such that Gr(zi)
~
Kl for of all
i == 1,2, ....
This contradicts the condition {(Zk)v: k = 1,2, ... } E s(V). Thus our assumption has led to a contradiction. The assertion is proved. Let us now see how results of this kind work. Example 3.1. Consider the system (3.1)
== -x + y + ).2 x - l e - l,\y/x: 1 y' == -x - y - ).2 x - l e -l.\y/x I.
X' {
•
224
CHAPTER 7
We put the last terms equal zero for x = o. The right hand sides of the equations contain the real parameter A E [0, (0). Let Ai --t 00 and Zi denote the solution space of the system (3.1) for A = Ai. Let a = {Zi: i = 1,2, ... }. Show that a E s(JR x JR2). I. Let V be a region in the space JR x JR2 of variables t, x, y, the closure [V] be compact and y do not meet the plane y = o. On the set [V] the last terms of the equations (3.1) converge uniformly to zero. Therefore there exists a number m > 0, such that J. Ix'(t)1 ~ m and ly'(t)1 ~ m for evx ery solution u = (x, y) E U{(Zi)V: i = 1,2, ... }. This implies that {(Zdv: i = 1,2, ... } E s(V). II. Let V be a region in the space JR x JR2 of variables t, x, y. t =const Let the closure [V] be compact and lie either in the half space Figure 7.2 JR x ((0, (0) x JR), or in the half space JR x (( -00,0) x JR). Consider the mapping cp : V --t JR, cp(t, x, y) = x + y (Figure 7.2). Apply Assertion 3.2 to the set E = {(t, x, y): t, x E JR, y = o}. By I and the estimate ••••• _ _ _ _ _ _ _ _ _ _ _ _ o
(3.2)
I(cp(t, u(t)))~1 ~ m
for some m > 0 we have {(Zi)V: i = 1,2, ... } E s(V). III. Regions mentioned in I and II cover the plane t = const except the point x = y = O. Now the membership a E s(U) follows from Corollary of Theorem 3.1. In results proved we deal not with the right hand side of an equation (inclusion), but we deal directly with a solution space. Therefore when we investigate an particular equation, we need preliminary considerations before we obtain any possibility of refering to these results. In order to obtain a corresponding result dealing with right hand sides of equations (inclusions), introduce the class of multi-valued mappings F : U --t ~n satisfying the conditions: (3.3) values of the mapping Fare nonempty convex one sided cones with the top at the origin of coordinates ({O'} may be the value too). Assume that:
225
Convergent sequences of solution spaces.
(3.4) every multi-valued mapping F, Fi : U condition (3.3);
--t
lRn ,
i = 1,2, ... , satisfies
(3.5) F(x) 2 n{ CC(U{Fi(OX) : i = k, k+1, ... }UF(Ox» : k = 1,2, ... , Ox runs over the set of all neighborhoods of x E U} for every point x E U; (3.6) the primitives of Lebesgue integrable functions
(}:i
i = 1,2, ... , constitute the equicontinuous family, Si(t)
Ilull
:
lR
--t
[0,(0), E lR,
= {u: u
~ (}:i(t)};
(3.7) Gi(t, y) = Fi(t, y) + Si(t) (t, y) E U and i = 1,2, ....
{u
+v
Multivalued mappings of these conditions will play the role of majorants of right hand sides in the investigation of particular equations and inclusions. Its geometry allows us to estimate the structure of sets of singularities. Their role is close to the role of the mapping () in Theorem 3.1. Its unique advantage with respect to the mapping () is its simple and close relation (majorization) with right hand sides of equations (inclusions) under consideration. For a cone V with the top at the origin of coordinates and c > 0 denote by Po the c-neighborhood of the subset p = {u: u E V, Ilull = I} of the unit sphere S and denote by Ve the cone
{O'} U
{u: uE lR ui= 0, n,
11:11
E Pc}
(see Figure 7.3).
/ -.... ,
Figure 7.3
Figure 7.4
Denote by r(F, t) the topology on the set Ut generated by the system
of neighborhoods {y ± (F(t, Y»c n 0 6 0: c,8 > O}, y E Ut (see Figure 7.4).
CHAPTER 7
226
Theorem 3.2. Let conditions (3.3 - 7) hold,
{Zi: i = 1,2, ... }
~
R(U),
Zi ~ D(G i ) for all i = 1,2, .... Let, be a family of open subsets of the set U. Let {(Zi)V: i = 1,2, ... } E s(V) for every V E ,. Let E = U\ (U,). For every t E ~ the set E t let contain no nontrivial continua on which the topology r(F, t) and the Euclidean topology coincide. Then {Zi: i = 1,2, ... } E s(U). ~
Proof. I. Assume the opposite. For a compactum K a sequence Zj
E (Zij)K,
il
U there exists
< i2 < ...
which is not equicontinuous. This means that for some c > 0 for every m = 1,2, ... the set Am of indices kEN satisfying the condition (3.8) there are points IIZk(t) - zk(s)11 ~ c,
S
< t of 7r(zm) such that t -
S
< 2- m and
is infinite. Since the formula
jl
= minAI,
ji
= min(Ai \ {il, ... ,ji-d) for i
= 2,3, ... ,
defines an infinite sequence of indices A = {i1,i2, ... }. For jm E A fix a pair of points Sjm < tjm according to (3.8): t·3Tn - s·J'fn
< 2- m and liz'3Tn (t·3m ) - z·]Tn (s·3Tn )11
2r c.
Thus (3.9)
tj -
Sj --+
0 as j E A and j
--+ 00,
(3.10)
The projection P of the compactum K in the first factor of the product IR x IRn is compact. Therefore the sequence {Sj : j E A} has a subsequence {Sj: j E B} converging to a point S E P. Due to (3.9) S = ~im tj. 1EB.
j-+oo
Convergent sequences of solution spaces.
The space exp K is compact, see Theorem 3.4.3. Therefore the sequence {Gr{zj): j E B} contains a subsequence {Gr{zj): j E Bd converging in exp K to a limit H. Since S
=
.lim
t--+ 00
Sj
=
227
!
/: ,, .
! i
.lim tj
1--+00
the projection of the compactum H into the first factor of the product IR x IRn coincides with the point s. Condition (3.10) implies that diam H ;;:: c. Therefore the set H contains more than one point. The compactum H is connected, see Theorem 3.3.3 and Figure 7.5. II. Show that
:
~
:
:
-s;------~--s;
--72
S3i3
S
Figure 7.5
H ~ U\ (U'Y).
Assume the opposite, i.e., the existence of a point (t, Yo) E H For some 'TJ > 0
(3.11)
H \ 0T/Yo
n (U,).
=1= 0,
(3.12) the set [OT/{t, Yo)] lies in an element V of the family 'Y.
By virtue of the condition (t, Yo) E H and Theorem 3.4.2 we can fix points Pj E [Sj, tj] in order that Zj{pJ -+ Yo as j E BI and j -+ 00. The last convergence implies the infiniteness of the set B2 = {j : j E B I , (Pj, Zj(pj)) E 0T/(t, Yo)}. For j E B2 denote by I j the maximal segment satisfying the conditions Pj E I j ~ [Sj,tj], Gr(zjlIJ ~ [OT/{t, Yo)]. Because of (3.11) the sequence {Gr(ZjIIJ:j E B2} has limit points outsides 0T/Yo, Thus the sequence {ZjII' : 'j E B 2 } is not equicontinuous. J This contradicts conditions (3.12) and {{Zi)V: i = 1,2, ... } E s{V). III. Assume that the continuum H t lies in a plane A ~ IRn. Let A denote the space of vectors of the plane A. {In the connection with the notation H t and with the analogous notation below recall that H ~ {t} x IRn and
228
CHAPTER 7
H t ~ ]Rn). If at all points of the set H t or of a nonempty open subset of it their above neighborhoods in Ut with respect to the topology reF, t) contain Euclidean 8-balls of the plane A, then the induced topologies coincide. For this case the theorem is proved. Analyze now the case when for a point Xl = (t, YI) E H the cone F(t, YI) does not contain a half-space of the space A. The set CC(U{Fi(Oxd: i = k, k + 1, ... } U F(Oxd) in (3.5) is a closed cone. We have two possibilities: A. The zero vector 0 of the space A does not belong to the interior of the set CC(U{Fi(Oxd: i = k, k + 1, ... } U F(Oxd) n A. B. The zero vector 0 of the space A belongs to the interior of the set
CC(U{Fi(Oxd: i = k,k
+ I, ... } UF(Oxd) nA.
Then CC(U{Fi(Oxd: i = k, k + I, ... } U F(Oxd) = A. If B is true for every neighborhood OXI of the point Xl and for every k = 1,2, ... then by virtue of (3.5) F(xd ;2 A, which leads out the framework of the case under consideration. Thus for a neighborhood OXI of the point Xl condition A holds. For some 1-£ > 0 we have [t - 1-£, t + I-£J x [OJ.lyd ~ OXI· Let el be an arbitrary point of the unit sphere of the space A, which does not belong to the cone
By Theorem 2.4.2 there exist a linear functional h : ]Rn ~ ]R and numbers p < q such that heel) ~ q and h(S) ~ (-oo,p). Since S is a cone, we can take p = o. There exists a vector eo E ]Rn, such that h( u) = (eo, u) for all u E ]Rn. For the vector e = -eo/lleoll we have: (e, u) ~ 0 for every vector u E CC(U{Fi([t - 1-£, t i = k, k + 1, ... } U F([t - 1-£, t + I-£J x [OJ.lyd)). (3.13)
+ I-£J
x [OJ.lYIJ):
IV. Return to the notation of I. Let a linear functional 9 : satisfy the condition
(3.13') g(u) ~ 0 for every vector u E CC(U{Fi([t - 1-£, t + 1, ... } U F([t - 1-£, t + I-£J x [OJ.lyd))·
+ I-£J
]Rn ~
lR
x [OJ.lYIJ)
i = k, k
In particular, by (3.13) this holds for the functional g(u) = (e, u). By Theorem 3.4.2 we can fix points rj E [Sj, tjl such that zj(rj) ~ YI as j E B 1 , j ~ 00. So the set B2 = {j: j E B 1 , r j E (t - 1-£, t + 1-£), zj(rj) E OJ.lYl} is infinite.
229
Convergent sequences of solution spaces.
Now let a set B3 ~ B2 be infinite. Let for j E B3 points Pj ~ qj of the segment [t - f.£, t + f.£] are such that Zj([Pj, qj]) ~ [OILYI]. Let CPj(t) = g(Zj(t)) for t E [pj, qilThe function CPj is generalized absolutely continuous and by (3.13') and
(3.7)
Prove that (3.15) if cPj(Pj)
-+
G, CPj(qj)
-+
D as j E B3 and j
-+ 00,
then G ~ D.
By (3.14) the function t
'ljJj(t) = CPj(t)
+
J
aj(s)ds
Pi
cPj(qj) - CPj(Pj)
~
qi
-
J
aj(s)ds
-+ 0
Pi by virtue of the equicontinuity of the primitives of the functions aj. Thus
This gives the fulfilment of (3.15). V. Keep the notation of III and IV. Assume that:
Gr(zj![Pi,qi]) -+HI ~ Hn({t} x (OILYI n>.))
as
j E B,
J -+ 00,
and that the mapping f(u) = (e, u) is injective on the compactum (Hdt. Show that the topology r(F, t) and the Euclidean topology induce on (HI)t the same topology. The continuity of the mapping f, the connectedness and the compactness of the set HI with respect to the Euclidean topology imply that the set f((Hdt) ~ IR is a segment (or an one point set). Denote it by I. By virtue of the compactness of the set (Hdt, of the continuity and of the bijectivity of the mapping f : (Hdt -+ I the inverse mapping
CHAPTER 7
230
9 : I -+ (Hdt is continuous with respect to the Euclidean topology on the set HI (Theorem 2.2.1). So it is sufficient to check the continuity of the mapping 9 with respect to the topology r(F, t). Assume the opposite. For a point (t, a) of the set HI and for a number c > 0 for every j = 1,2,... there exists a point aj E ). satisfying the condition
By virtue of the compactness of the unit ball we can pass to a subsequence
of the sequence
{ II:; =:11 : j
= 1,2, ... }
converging to a vector Uo E A. We have lIuoll = 1 and Uo (j. ±(F(t,a)hc. Conditions (3.3) and (3.5) and Lemma 2.5.2 imply that
(F(t, a»)e: 2 CC(U{Fi(O(t, a»: i = k, k + I, ... }) for some k = 1,2, ... and for some neighborhood O(t, a) of the point (t, a) in U. By virtue of the continuity of the mapping 9 with respect to the Euclidean topology g([f(a) - 8, f(a) + 8]) ~ O(t, a) for some 8 > O. The vector Uo does not belong to the set
Lo
= CC(U{Fi(O(t, a)):
i
= k, k + 1, ... })
nor to the set
Ll
=-
CC(U{Fi(O(t, a»: i
= k, k + 1, ... }).
Theorem 2.4.2 and the membership 0 E Lo n Ll imply the existence of linear functionals fi : JRn -+ JR, i = 1,2, satisfying the condition
By virtue of the convergence aj -+ a with respect to the Euclidean topology and the continuity of the functional f we have f(aj) -+ f(a). Thus f(aj) E [f(a) - 8, f(a) + 81 beginning with some j = k.
Convergent sequences of solution spaces.
231
The segment Mj with endpoints f(aj) and f(a) lies in the segment + 8]. Therefore
[f(a) - 8, f(a) (3.16)
We obtain: If(aj) - f(a)1 < 8, aj E O(t, a) and g(Mj ) ~ O(t, a). By virtue of the continuity of the functional Ii, i = 1,2, we have fi ((a - ai)/ilai - all) < 0 beginning with some j = kl ~ k. Therefore for j ~ kl (3.17) Nw take an arbitrary index j ~ k l . Since the mapping f is injective, f(ai) =I f(a). Consider the case when f(ai) > f(a). By Theorem 3.4.2 for j E B3 we can fix points ej, dj E [pj, qj] such that
Zj(Cj)
-t
a,
zj(dj ) - t ai.
The continuum HI is homeomorphic to a segment. By (3.16) between the points a and ak it does leave the set O(t, a). Then the graph of the function Zj between Cj and dj lies in set O(t, a), beginning with some j = jo E B 3 • By (3.15) Cj < dj , beginning with some j E B 3 • The latter fact, the inclusions Gr (Zjl[Cj,d j]) C;;; O(t,a), fo(Lo) C;;; [0,00), and the considerations of IV with 9 = fo imply that fo(a) ~ fO(ai), which contradicts (3.17). The case f(ai) < f(a) may be considered in an analogous way with the change fo to fl. Thus our assumption is false. Hence the mapping 9 is continuous with respect to the topology r(F, t). In the beginning of our reasoning we noticed that it implies the coincidence on the set HI of the topology r(F, t) and of the Euclidean topology. VI. Return to the notation of III and put for j E B2
Pj = inf{p: p E [Sj, Tj], p ~ t - J.L, Zj([P, TjD
C;;; [O~yd},
qj = sup{q: q E [Tj, tj], q ~ t + J.L, Zj([Tj, qD
C;;; [O~yd}·
The set HI from V contains more than one point. If the corresponding mapping f is injective we obtain what was required from V. In the opposite case for some mE f«Hdt) the set H2 = ({y} xf-l(m))nHl contains more than one point. Let a and b be its arbitrary two different points. According to Lemma 3.4.2 fix points Pj, qj E [pj, qj] in order to obtain the convergence Zj(Pj) - t a, Zj(qj) - t b.
232
CHAPTER 7
Assume for definiteness (with a passage to a subsequence, if it is necessary) that pj < qj. Let L = ]-1(0) n A. If for every (t, y) E H2 the set F(t, y) contains entirely the space L then the topology r(F, t) coincides on (H2 )t with the Euclidean topology. In the opposite case we return to the beginning of the reasoning of VI with the change A to An ]-l(m), Pj, qj to Pj, qj and we do it till the fulfilment of assumptions of V. Since in every step the dimension of the corresponding plane A decrease by 1, not late than in n-th step our process finishes. We cannot come down to the dimension zero, because in this case a nontrivial continuum would be a subset of the plane of the dimension zero. The theorem is proved. • Example 3.2. Return to the system (3.1). Our aim here is the same as in Example 3.1. We repeat the steps I and III of the reasoning of Example 3.1 word by word. In the step II we have a possibility to use Theorem 3.2 with Si(t) == m (where m is from (3.2) ) and Fi(t, y) = {(A, -A): A ~ O}, that gives what was required: with the notation of Example 3.2 the sequence of solution spaces {Zi: i = 1,2, ... } belongs to s(JR. x JR.2).
4. Estimates of set location Let us go on to the second of the questions stated in the beginning of the previous section. Here we make some preliminary considerations. As before U denotes an open subset of the product JR. x JR.n. Let M ~ U, Mo ~ JR. and 0 =1= 4> ~ Cs(U). The upper bound j.£} is called a measure of the set M with respect to the space 4>. For rp E 4> denote by l(Mo, rp) the set of all connected open subsets ofthe segment 7r( rp) meeting the set Mo. For e > 0 denote by 1* (Mo, rp, e) the set of all finite pairwise disjoint families 'Y of elements of l(Mo, rp) with db) ~ e. If the set l*(Mo, rp, e) is empty we put w;,o(Mo) = o. In the opposite case we put w;,o(Mo) = sup{E{lIrp(sup J) - rp(inf 1)11 : J E A}: A E l*(Mo, rp, e)}. Notice that w* e > o. Therefore the limit w;(Mo) = limo--+o w;,o(Mo) exists. Let w
}. The set M is called rp-w-thin if the set {t: t E 7r(rp) , (t, rp(t)) E M} may be represented as a union of a countable number of its subsets M k , k = 1,2, ... , with w;(Mk ) = o. The set M is called 4>-w-thin if it is rp-w-thin for every function cp E 4>. The set M is called cp-thin (respectively, 4>-thin)
Convergent sequences of solution spaces.
233
if /-,
h(t, z{t» - h{s, z{s»
~
c{t - s) for every two points s < t of 7r{z).
Lemma 4.2. Under the imposed conditions the set Ml(h, c) is closed in the space Cs{U). Proof. Let a sequence {Zk: k = 1,2, ... } ~ Ml{h,c) converge to a function Z E Cs(U) and points s < t belong to 7r{z). By Theorem 3.4.2 we can choose points Sk, tk E 7r{Zk) such that (Sk' Zk{Sk» -+ (s, z(s» and (tk' Zk(t k ) -+ (t, z(t)). By virtue of the convergences Sk -+ sand tk -+ t we have Sk < t k, beginning with some k (see Lemmas 2.1.1 and 1.7.2). Since Zk E Ml(h,c),
By virtue of the continuity of the function h and Lemma 1.7.2 the limit passage in the last inequality implies that
h{t, z{t» - h(s, z(s»
~
c(t - s).
E Ml (h, c). The lemma is proved. • Lemma 4.3. Let a mapping h : U -+ 1R be continuous, c> 0, M ~ U and /-,(h{M» = O. Then /-'M1(h,c){M) = O. Proof. Take an arbitrary function Z E Ml (h, c) and put Mo = {t : t E 7r(z), (t, z{t» EM}. The function cp{t) = h(t, z(t» is defined on the segment 7r(z). It is continuous and increasing. Therefore it maps bijectively the segment 7r(z) onto a segment I (see Theorems 1.7.8, 2.6.3 and Corollary 1 of Theorem 2.6.4). By Theorem 2.2.1 the inverse mapping ¢: 1-+ 7r(z) is continuous. The membership Z E Ml (h, c) implies that the mapping'I/J satisfies the Lipschitz condition with the constant c- l . Therefore it is absolutely continuous (see Example 4.5.1). By Corollary of Theorem 4.5.1 the set Mo ~ 'I/J(I n h(M» has the measure zero. The lemma is proved. • Theorem 4.1. Let a mapping h : U -+ 1R be continuous, c> 0, M ~ U, . U(h(M» = 0 and Z ~ M 1 (h, c). Then JL[zj(M) = O.
So
Z
CHAPTER 7
234
Proof. Lemma 4.2 implies the inclusion [Z] ~ Ml (h, c). Referring to Lemma 4.3 completes the proof. • Let as in the previous reasoning a function h : U -+ IR be continuous. Let c ~ O. Denote by M 2(h, c) the set of all functions Z E Cs(U) satisfying the condition
Ilz(t) - z(s)11 7r(z).
~
c(h(t, z(t)) - h(s, z(s))) for every two points s < t of
Analogously Lemma 4.2 we prove: Lemma 4.4. Under the imposed conditions the set M 2(h, c) is closed in the space Cs(U). • Lemma 4.5. Let a mapping h : U -+ IR be continuous, c > 0, M be a closed subset of the set U and {L(h(M)) = O. Then WM2 (h,c)(M) = O. Proof. Take an arbitrary function z E M2 (h, c) and put Mo = {t : t E 7r(z), (t, z(t)) EM}, cp(t) = h(t, z(t)). We have cp(Mo) ~ h(M) and {L(h(M)) = O. By Lemma 4.1.1 {L(cp(Mo)) = O. Therefore for every c > 0 there exists an at most countable family al of interval such that Ual 2 Mo and d(al) ~ c-1c. The set a2 of connected components of the (open) set G = Ual covers Mo. By Theorem 4.1.2 {L(G) ~ c-1c. Elements of the set a2 are pairwise disjoint. By Theorem 2.6.7, Lemma 4.1.3, and Theorem 4.1.1 d(a2) ~ c-1c. The family {HA = Mo n cp-l(A): A E ad consists of pairwise disjoint sets. The closedness of the set M implies the closedness of the set Mo ~ 7r(z) (see Theorems 2.1.1 and 1.7.5). By Theorem 1.6.1 the set Mo ~ 7r(z) is compact. By the definition of the induced topology the set H A , A E a2, is open in the compactum Mo. The representation HA = Mo \ (U{HB: B E a2 \ {A}}) implies now its closedness. Elements of the family {HA : A E a2, HA =1= 0} are pairwise disjoint and open (in Mo). The compactness of the set Mo implies the finiteness of this family. Therefore the number 1/ = min{p(HA' 7r(z) \ cp-l(A)): A E a2, HA =1= 0} is positive. The membership z E M 2 (h, c) implies that the function cp is non-decreasing. For, E l*(Mo, z, 1/) we have
l:{llz(b)-z(a)1I : (a, b) E ,} ~ cl:{lcp(b)-cp(a)1 : (a, b) E ,} ~ c(c-1c) = c. So wz(M) = O. The last estimate gives immediately what was required. The lemma is proved. • Theorem 4.2. Let a mapping h : U -+ IR be continuous, c > 0, M be a closed subset of the set U, {L(h(M)) = 0 and Z ~ M 2(h, c). Then w[zj(M) = O. Proof. Lemma 4.4 implies the inclusion [Z] ~ M 2 (h, c). By Lemma 4.5 this gives what was required. The theorem is proved. •
Convergent sequences of solution spaces.
235
We say that a set M ~ U is at most countable (respectively, finite) with respect to a space Z ~ Cs(U) if for every function z E Z the set {t: t E 7l"(z), (t, z(t») EM} is at most countable (respectively, finite). The remarks of Example 4.1.1 imply: Theorem 4.3. Let Z ~ Cs(U). Let a set M ~ U be at most countable • with respect to Z. Then I-Lz(M) = o. The following assertion is obvious: Lemma 4.6. If z E Cs(U) and a set M ~ U is represented as a union of (an at most countable number of) its z-w-thin subsets M k , k = 1,2, ... , then the set M is z-w-thin. • Evidently an one point set M ~ U is z-w-thin. This fact and Lemma 4.6 imply: Theorem 4.4. Let Z ~ Cs(U). Let a set M ~ U be at most countable • with respect to Z. Then the set M is Z -w-thin.
5. Estimates of the upper limit of a space sequence As before U denotes an open subset of the product IR x IRn. Lemma 5.1. Let a family "I of open connected subsets of a segment [a, b], a < b, of the real line satisfy the conditions a) if VI, V2 E"I and VI n V2 =F 0, then VI U V2 E "I; b) for every nonempty closed subset F of the segment [a, b] there exists a (nonempty) set V E "I, which meets the set F. Then [a, b] E "f. Proof. I. The set G = U"I is open (in the segment [a, b]). Therefore the set F = [a, b] \ G is closed. Condition b) implies the emptity of the set F, i.e., the equality G = U"f. II. The set H = U"Io, where "10 = {V: V E "I, V 3 a}, is open in the segment [a, b]. Theorem 2.6.1 implies the connectedness of the set H. By Theorem 2.6.3 the set H either coincides with the segment [a, b], or is a half interval [a, c) for some c E (a, b]. In the second case I implies the existence of a set VI E "I containing the point c. The equality [H] = [a, c] implies the existence of a point t E H n VI. The definition of the set H implies the existence of a set Vo E "10 containing the point t. By condition a) we have Va U VI E "I. Since a E Vo ~ Va U VI, Va U Vi E "10. Hence c E VI ~ Va U VI ~ H. This contradicts the definition of the point c. In the first case there exists an element Vi of the family "11 containing the point b. So Vi = [a, b]. The lemma is proved. • Theorem 5.1. Let ~ E Ri(U) (see §6.2), Z E Rq(U), "I be a family of open subsets of the set U. Let ~v ~ Zv for every V E "I. Let the set U \ (U"I) be ~-thin. Then ~ ~ Z.
236
CHAPTER 7
Proof. I. Take an arbitrary function cP E 0 we have 'IjJ(Oot) ~ V. By virtue of the conditions O. The segment I = [t - 8, t + 8] n 11"( cp) may be represented as the union of a countable number of segments 10 = {t}, Ik = In (-00, t - 2k] for k = -1,-2, ... and h = In[t+2- k ,(0) for k = 1,2, .... For every k = ±1, ±2, ... the segment h does not meet the set Mo. Thus it is covered by the family v* = {h n G: G E vo}. This family and the segment h satisfy the condition b) of Lemma 5.1. By Lemma 4.9.1 condition a) holds too. Lemma 5.1 implies now the generalized absolute continuity of the function cplh•. The function cplIo is also generalized absolutely continuous. This implies the generalized absolute continuity of the function cplI. So I E v. The membership (t E) (I) E Vo contradicts the choice of the point t: t E Mo. This gives what was required. III. Show that UVo = 1I"(cp), i.e., that Mo = 0. By virtue of the cp-w-thinness of the set U \ (U,) and Lemma 4.1 the closed subset Mo of the segment 11"( cp) may be represented as the union of an at most countable family {Mi: i = 1,2, ... } with w;(Mi ) = O. By virtue of the definition of the induced topology Baire theorem 1.7.6 implies the existence of points c < d of the segment 1I"(cp) and of a number i = 1,2, ... such that 0 =1= Mo n (c, d) ~ [Mi] n (c, d). Denote M; = Mo n (c, d). For arbitrary 'f} > 0 there exists 8 > 0 such that w;,o(Mi ) < 'f}. Let a family r satisfy condition (*,8, M;) of §4.5. Enumerate elements of the family r according to its order on the line: r = {(ai, bi ): i = 1, ... , k}, where a1 < b1 :::; a2 < ... :::; ak < bk · Put 101
= min( {bi -ai : i = 1, ... ,k }U{ ai+l -bi
:
i
= 1, ... ,k-1, aH1 -bi
=1=
O}).
Let £ E (0,101/2) and 8 E U{{ai,b i }: i = 1, ... ,k}. The point 8 E M; is not isolated in the set Mo ~ [Mil. Thus the set (8 - £,8 + c) n Mi
Convergent sequences of solution spaces.
237
is infinite. Therefore we can fix a point SE of this set such that the sets (8 - c, SE) n Mi and (SE, S + c) n Mi are nonempty. The sets (s - c, s + c), where s E U{ {ai, b;}: i = 1, ... ,k}, are pairwise disjoint. Thus the family fe = {((ait, (bit): i = 1, ... ,k} consists of pairwise disjoint (nonempty) intervals. For c < 1/2k(o - E{b i - ai: i = 1, ... ,k}) we have d(f':) < o. Because of the choice of the points SE every element of the family re meets the set Mi' Thus
E{lIcp((a;)E) - cp((bit)11 : i = 1, ... , k} ~ 'f/. By virtue of the continuity of the function cp the passage to the limit in this inequality as c ---+ 0 gives the estimate
This proves the absolute continuity of the function cp on the set M;. By Theorem 1.6.10 the family Vl = {I: [E vo, [ ~ ((c, d) n1f(cp)) \M;} contains an at most countable subfamily {Ik: k = 1,2, ... }, the union of which coincides with UV1' Lemmas 5.1 and 4.9.1 imply the generalized absolute continuity of the restrictions of the function cp on every segment [Ie, k = 1,2, .... Since (c, d) n 1f(cp) = M; U U{[k: k = 1,2, ... } we obtain the generalized absolute continuity of the restriction of the function cp to the set (c, d) n 1f(cp). Next we obtain the emptity of the set Mo. IV. Lemmas 5.1 and 4.9.1 imply now the generalized absolute continuity of the function cpo V. Let a segment [ lie in 1f(cp) \ M. Denote by e the set of all connected open subsets G of the segment [satisfying the condition: cplJ E Z for every segment J ~ G. The condition Z E R(U) implies, that the set e satisfies condition a) of Lemma 5.1. Our choice of [ and the definition of M imply that e satisfies condition b). By Lemma 5.1 CPII E Z. Since J.L(M) = J.L",(U \ (U,)) = 0 and Z E Rq(U), cp E Z. In view of the arbitrariness of cp E we have ~ Z. The theorem is proved. • Lemma 5.2. Let {Zk: k = 1,2, ... } ~ Ri(U) and Z = lim tOPSUPZk' Then Z E Ri(U). k->oo Proof. Let z E Z. Let a segment [a, bJ lie in the domain [s, tJ of a function z. Select a sequence Zk E Zk, k E A, converging to the function z (see Lemma 1.5.1). Let 1f(Zk) = [Sk' tkJ. By Corollary of Theorem 3.5.1 Sle --.. sand tk ---+ t. Let Sk (respectively, t k ) be a point of the segment [Sk' tkJ nearest to the point a (respectively, to the point b). We have {a, b} ~ [s, tJ, Sic --.. sand tk ---+ t. Therefore Sk ---+ a and tk ---+ b. In addition st, ~ tt, and [sk' tkJ ~ 1f(Zk)' By Theorem 3.5.4 zlcl[s;,t;) ---+ Zl[a,b)' The lemma is proved. •
238
CHAPTER 7
When we apply the just proved lemma to a stationary sequence Z k == Z, we obtain: Corollary. Let Z E Ri(U). Then [Z] E Ri(U). • Theorem 5.2. Let Z E Rq(U). Let 'Y be a family of open subsets of the set U. Let the set Zv be closed in Cs(V) for every V E 'Y. Let the set U \ (U'Y) be [Z]-thin. Then the set Z is closed in the space Cs(U). Proof. By Corollary of Lemma 5.2 the space cI> = [Z] belongs to Ri(U). By our hypotheses cI>v = Zv for every element V of the family 'Y. Theorem 5.1 implies the inclusion cI> ~ Z. The inverse inclusion Z ~ cI> is obvious. The theorem is proved. • Notice that the application of this theorem may remain valuable and convenient when the family 'Y covers the entire set U. In this case we have: Corollary. Assume that Z E Rq(U), 'Y is an open cover of the set U for every element V of which the set Zv is closed in the space Cs(V). Then the set Z is closed in the space Cs(U), • We will reinforce this corollary in §6 and we will change the condition Z E Rq(U) to the condition Z E R(U). In respect of other applications of Theorem 5.2 pay attention to Theorems 4.2-6 wich help to obtain the estimates occuring in Theorem 5.2. Theorem 5.3. Let Z E Rq(U), 0 = {Zk: k = 1,2, ... } ~ Ri(U), cI> = lim top sup 0 ~ w. Let 'Y be a family of open subsets of the set U. Let lim top SUPk-> 00 (Zk)V ~ Zv (in Cs(V») for every V E 'Y. Let the set U \ (U,) be w-thin. Then cI> ~ Z. Proof. By Lemma 4.1 the set U \ (U'Y) is cI>-thin. Referring to Theorem 5.2 completes the proof. • By Lemma 1.4 Theorem 5.3 implies: Theorem 5.4. Let under hypotheses of Theorem 5.3 a E s(U). Then the sequence a converges to the space Z in U. • Theorems of this section in their full generality are directed to the investigation of singularities (ofright hand sides with parameters), including rather complicated ones. Now, in order not to cause the reader to digress from the scheme of application of results of such a kind by technical difficulties, consider examples of the investigation of the simplest singularities. Example 5.1. The right hand side of the scalar equation (5.1 )
,
y =
0 0 2
+ t2 + y2
is defined for all real values of 0, t and y, except (the one point of the three-dimensional space ]R3) 0 = t = Y = O. Define the right hand side there by putting its equal to zero for this values of 0, t and y. By remarks of Example 6.5.2 for every value of the parameter 0 the solution space of our equation satisfies conditions (c), (e) and (110). By remarks of §1 solutions of
Convergent sequences of solution spaces.
239
the equation depend continuously on the parameter a for a #- o. Consider an arbitrary sequence ak ~ 0, ak i- 0, and denote by Zk the solution space of our equation with a = ak. Denote by Z the solution space of the equation with a = O. For every point (t, y) i- (0,0) of the plane ~2 denote by V(t, y) the neighborhood of the point (t, y) of the radius tv't2 + y2. Put 'Y == {V(t, y): t, y E ~2, (t, y) i- (0, By remarks of §1 for every V E , the sequence {(Zk)V: k = 1,2, ... } converges to the space Zv in V. The set E = ~2 \ (U,) consists of one point (0,0). By Remark 1.1 and Corollary 2 of Theorem 3.1 the sequence 'fJ = {Zk: k = 1,2, ... } belongs to S(~2). By Theorems 4.3 and 4.4 the set E is 'IT-thin for 'IT = C8(~2). By Theorem 5.4 the sequence 'fJ converges to the space Z in ~2. The latter fact, in view of the arbitrariness of the sequence ak ~ 0 and Theorem 2.5.1, means the continuity of the dependence of solutions of our equation on the parameter a for a = O. Thus solutions of our equation depend continuously on the parameter a (for every its value). In a completely analoguous way we can consider the equation
On.
(5.2) where the function f : IR x IR ~ ~ is continuous. Example 5.2. Let a function f : ~ x ~ ~ [2, 00) be continuous, a E R Show that solutions of the equation (5.3)
y' = f(t, y)
+
a2
lal
+y2
depend continuously on the parameter a (as in the previous example for a = 0 we put the fraction equal to zero independently on the value of V). For a i- 0 the right hand side of equation (5.3) is continuous in the totality of the variables a, t, y. Here we obtain what was required from Theorem 1.3. Thus it remains to prove that for every sequence ai ~ 0, ai i- 0, the sequence of the solution spaces of the equations
(5.4) converges to the space Z of the solutions of the equation y' = f(t, V). Outside the line y = 0 the right hand side of the equation (5.3) depends continuously on the totality of the variables a, t, y. By Theorem 1.3 (Zi)V ~ Zv as i ~ 00, where V = ~x (~\ {O}) and Zi denotes the solution space of the equation (5.4). By Remark 1.1 and Corollary 1 of Theorem 3.1 the sequence of the spaces {Zi: i = 1,2, ... } belongs to s(~ x ~). To complete the reasoning
CHAPTER 7
240
we need to obtain estimates of the set (~ x ~) \ V = ~ x {O} according to considerations of §4. By putting h(t, y) = y and c = 1 we obtain the estimates JlM,(h,c)(E) = WM2(h,c)(E) = O. The solution spaces of all equations under consideration lie in the closed subspace W = Ml (h, c) n M 2 (h, c) of the space Cs (~ x ~) Now Theorem 5.4 implies the convergence Zi ~ Z. We obtain what was required. Results of the previous section and last examples prepare us for the consideration of the question about the convergence Zi ~ Z in steps: we can prove first the convergence (Zi)V ~ Zv for an arbitrary element V of a family of open subsets of the set U, and then pass to estimates of the geometry of the rest U \ U,. The convergence (Zi)V ~ Zv may be proved, for instance, with the use of classical theorems or with the use of any new versions of the corresponding classical theorems. The possibility of proving the convergence (Zi)V ~ Zv with the use of our scheme itself following our estimates of the geometry of the singularity set is not very obvious. But we can repeat the estimating of the geometry of singularity set several times, even we can use here a transfinite induction. Example 5.3. Keep the assumptions of Example 5.2. Consider the equation
y' = f(t, y)
(5.5)
+
lal
a 2 + y2
+
lal
a 2 + (y - sint)2
+
lal
a 2 + (y - cost)2
As in Examples 5.1-2 solutions of equation (5.5) depend continuously on the parameter a (as in two previous examples for a = 0 we put the fraction equal to zero independently on the values of t and y). For a i- 0 the right hand side of the equation (5.3) is continuous in the totality of the variables a, t, y. We obtain what was required from Theorem 1.3. In order to analyze the situation for a = 0 consider an arbitrary sequence ai ~ 0, ai i- O. Let Zi denote the solution space of the equation
y
, = f(t
,y
)+
lail
ai 2 + y2
+
lail
ai 2 + (y - sin t)2
+
lail
ai 2 + (y - cos t)2
Let W denote the complement in the plane ~ x ~ to the union of the x-axis L 1 , of the graph L2 of the function y = sin t and of the graph L3 of the function y = cos t. The convergence (Zi)W ~ Zw as i ~ 00 follows froIll Theorem 1.3. Every point x of the set E = (~ x ~) \ W belonging to one only of the sets L 1 , L2 or L3 possesses a neighborhood Ox without points of the other two of these three sets. The convergence (Zi)Oo: ~ Zoo: as i ~ 00 may be proved as was done in Example 5.2.
Convergent sequences of solution spaces.
241
Figure 7.6
The set (L1 n L 2) U (L2 n L 3 ) u (L1 n L 3 ) (Figure 7.6) consists of isolated points. We have the result of the previous paragraph. Now we repeat the reasoning of Example 5.1 which gives the convergence Zi -+ Z. We obtain what was required. Example 5.4. Return to system (3.1). Denote by Z the solution space of the system (5.6)
{
Xi
= -x
yl =
+y
-x - y.
Let U = JR X (JR 2 \ {O}). Show that with the notation of Example 3.1 the sequence ao = {(Z;)u : i = 1,2, ... } converges to Zu. I. With the notation of step I of Example 3.1 (5.7)
(Zi)V
-+
Zv·
This follows from classical results (see §1). II. Let e > 0, V = JR x ((2e,3e) x (-e,e)), h(t,x,y) = -y and W= M1(h,e). The spaces (Zl)V, (Z2)V, (Z3)V,'" lie in w. By I it remains to investigate the set E = V n (JR x (JR x {o})). It is finite with respect to w. By Theorem 5.7 we have (5.7). III. If e > 0 and V = JR x ((-3.::, -2.::) x (-.::, .::)), then we have (5.7) This is analogous to II. IV. The regions V studied in I-III cover the set U entirely. By Theorem 5.7 we have what was required. The origin of coordinates 0 is an asymptotically stable stationary point ofthe system (5.6). The relation between the systems (3.1) and (5.6), which is lllentioned in Example 5.4, is sufficient to state that the point 0 is also an asymptotically stable stationary point of the system
242
CijAPTER 7
Here the usual effects of 'first approximation' keep their place, although in our case in the x-axis (Le., for y = 0) perturbation terms can be written as ~ and they tend to the infinity as x ---+ O. This lead us far from common conditions of the 'first approximation', see Chapter XV. 6. Additional remarks
As before U denotes an open subset of the product ~ x ~n. Theorem 6.1. Let cP E Ri(U), Z E Rp(U)" be a family of open subsets of the set U, for every element V of which CPv ~ Zv. Let the set U \ (U,) be at most countable with respect to CP. Then cP ~ Z. Proof. Take an arbitrary function cp E CPo Denote by /J the set of all segments I ~ 7r(cp) such that cpll E Z. Put /Jo = {(I): IE /J}, M = {t : t E 7r(cp), (t, cp(t)) E U \ (U,)}. The set M is closed, at most countable and 7r(cp) \ (U/Jo) ~ M. The family /J satisfies obviously condition a) of Lemma 5.1. Show that it satisfies condition b). It is sufficient to consider the case when F ~ M (in the opposite case the corresponding condition is satisfied obviously). Then the set F is at most countable. By Assertion 2.4.1 there exists an isolated point to of this set. Take 8 > 0 such that 02cton(F\ {to}) = 0. The family /Jo covers the segments II = [to - 8, to] n 7r(cp) and 12 = [to, to + 8] n 7r(cp) except its endpoint to. The condition Z E Rp(U) implies that cplll' cplh E Z and cplltul2 E Z. So we have shown the fulfilment of condition b). By Lemma 4.6 cp E Z. The theorem is proved. • Theorem 6.1 implies easily (see the proof of Theorem 5.4) Theorem 6.2. Let a = {Zk: k = 1,2, ... } ~ Ri(U), a E s(U), lim top SUPk-+oo Zk ~ \If ~ Cs(U), Z E Rp(U). Let, be a family of open subsets of the set U. Let the sequence {(Zk)V: k = 1,2, ... } converge in V to the space Zv for every V E ,. Let the set U \ (U,) be at most countable with respect to \If. Then the sequence a converges in U to the space Z. • This assertion remains valuable in the case when the family, covers the entire set U although here is better to use the following simplified analog of Theorem 6.1. It is a direct consequence of Lemma 5.1 (see the proof of Theorem 6.1): Theorem 6.3. Let cP E Ri(U), Z E R(U), , be an open cover of the • set U, for every element V of which CPv ~ Zv. Then cP ~ Z. Corollary. Let Z E R(U), , be an open cover of the set U, for every element V of which the set Zv is closed in the space C s(V). Then the space Z closed in C 8 (U). • As an other consequence of this theorem we obtain: Theorem 6.4. Let a = {Zk: k = 1,2, ... } ~ Ri(U), Z E R(U). Let , be an open cover of the set U. Let the sequence {(Zk)V: k = 1,2, ... }
Convergent sequences of solution spaces.
243
converge in V to Zv. for every V E 'Y. Then the sequence a converges in
U to Z. Proof. The proof repeats the proof of Theorems 5.4 and 6.2 with the use of referring to Theorem 6.3. The fulfilment of the condition a E s(U) follows here from Corollary of Theorem 3.1 and from the emptity of the set U \ (U'Y). • Remark 6.1. The convergence of a stationary sequence Zk == Z as k -+ 00 in U to the space Z E R(U) is equipotent to the condition
Z E Rc(U). Therefore Theorem 6.4 implies: Corollary. Let Z E R(U). Let'Y be an open cover of the set U for every element V of which Zv E Rc(V). Then Z E Rc(U), • Likewise from Theorems 5.4 and 6.2 we obtain the following assertions. Theorem 6.5. Let Z E Rqs(U), lJI ;2 Z be a closed subspace of the space Cs(U), Let'Y be a family of open subsets of the set U. Let Zv E Rc(V) for every V E 'Y. Let the set U \ (U'Y) be lJI-thin. Then Z E Rc(U). • Theorem 6.6. Let Z E Rps(U), lJI ;2 Z be a closed subspace of the space Cs(U). Let'Y be a family of open subsets of the set U. Let Zv E Rc(V) for every V E 'Y. Let the set U \ (U'Y) be at most countable with respect to lJI. Then Z E Rc(U). • The following particular case of Theorem 3.1 may be helpful to verify the fulfilment of the condition Z E Rs(U) when we use Theorems 6.5 and 6.6. Theorem 6.7. Let Z E R(U), () be a multi-valued mapping of U in ~n. For every point (t, y) E U let lim topsup{Gr(zl[Sl,S2]): z E Z, [SI' S2] ~ (t - c, t
+ c n 7r(Z)),
6,0->0
Z([SI' S2])
n G6y i= 0}
~ (}(t,
y).
Let 'Y be a family of open subsets of the set U. Let Zv E Rs (V) for every V E 'Y. For every point t E ~ let the set (U \ (U'Y))t contain no nontrivial connected subsets on which the topology r((), t) and the Euclidean topology coincide. Then Z E Rs(U). • As in the case of Theorem 3.1, we obtain: Corollary. Let Z E R(U). Let'Y be a family of open subsets of the set U, for every element V of which Zv E Rs(V). For every point t E ~ let the 8et (U \ (U'Y))t contain no nontrivial connected subsets. Then Z E Rs(U). •
7. R(U) as a topological space As before U denotes an open subset of the product ~ x ~n.
244
CHAPTER 7
For M ~ U and V ~ C.(U) put O{M, V} = {Z: Z E R(U), ZM ~ V}. Introduce the topology on the set R(U) generated by the sub-base consisting of the sets O{M, V}, where M runs over the set of all compact subsets of U and V runs over the set of all open subsets of the metric space C.(U). Notice that the topology introduced is not Hausdorff, but this fact is not an obstacle to its using. Lemma 7.1. Let Z E Rc(U). Then the point Z of the space R(U) possesses a countable base. Proof. I. Fix a countable base (3 of the set U. Denote by (30 the set of elements of (3 with compact closures lying in U. The set (30 is countable and constitutes a base. Theorem 1.1. 7 implies that the set (31 of all finite subsets of the set (30 is countable. Thus the family x = {[U,]: , E (3d is countable. It consists of compacta lying in U. By Theorem 1.1. 7 this implies the countability of the family 8 = {O{K, O(ZK' 2-i)}: K E x, i=1,2, ... }. II. Take an arbitrary compactum K ~ U and an arbitrary open subset V :2 ZK of the space C.(U). Show that for some c > 0 we have O,(K, c) ~ U and ZO,(K,E) ~ V. Assume the opposite. For i = 1,2, ... and Ci = 2-ip(K, (~ x ~n) \ U), there exists a function Zi E ZO,(K,Ei) \ V. The set ZO,(K,Ed is compact. It contains all elements of the sequence a = {Zi: i = 1,2, ... }. Therefore the sequence a has a subsequence converging to some Z E Z. Then Z E n{ZO,(K,E;}: i = 1,2, ... } = ZK and Z (j. V. This is impossible by the choice of K and V. III. With the notation of II for every point x E K fix an element Bx of the base (30 such that x E Bx ~ O( K, c). The cover {Bx: x E K} of the compactum K contains a finite sub cover {B X1 ' ... ' BXj }. Put K1 = [U{Bxw .. ,Bxj }]. Evidently K1 E x and K ~ K1 ~ O,(K, c). By virtue of the last inclusion ZK ~ ZK 1 ~ V. By the Corollary of Lemma 2.3.8 O(ZKl' 2- i ) ~ V for some i = 1,2, .... Therefore
The set O{K1' O(ZKIl2-i)} belongs to 8. By virtue of the arbitrariness in the choice of K and V the inclusion Z E O{K1' O(ZKIl2-i)} ~ O{K, V} means that the family 8 constitutes a sub-base at the point Z. The family of intersections of finite subfamilies of 8 is countable and constitutes a base at the point Z. The lemma is proved. • Lemma 7.2. A sequence a = {Zi: i = 1,2, ... } ~ R(U) converges to Z E Rc(U) in the space R(U) if and only if it converges to Z in U in the sense of the definition of §1.
Convergent sequences of solution spaces.
245
Proof. By virtue of Lemma 1.7.3 the convergence of the sequence a to Z in the space R(U) is equipotent to the condition (7.1) for every K E expc U the sequence {(Zd K : i = 1,2, ... } converges in the space Cs{U) to the set ZK in the sense of condition (1.7.1).
By Lemma 2.5.1 condition (7.1) is equipotent to the convergence of the sequence a in U to the space Z in the sense of §1. The lemma is proved .• Now Lemmas 7.1 and 7.2 and Theorem 1.5.3 imply: Theorem 7.1. A point Z E Rc{U) belongs to the closure of a set M ~ R{U) if and only if there exists a sequence of elements of the set M converging in U to the space Z. • So tools developed in previous sections of this chapter are applicable to . the description of the topology of the subspace Rc{U) of the space R{U). On the other hand, the same relation of the convergence of sequences with the topology of the space Rc{U) simplifies the using of results about the preservation of properties of solution spaces under limit passages. For instance, Theorem 2.1 states the closedness of the set Rce{U) in the space Rc{U). When they establish properties of solution spaces of equations (inclusions) often it is convenient to construct corresponding approximating sequences of spaces. In order to use Theorem 2.1 to prove the fulfilment of condition (e), we may construct a space sequence converging to the space in question (and satisfying the other hypotheses of Theorem 2.1). We will meet such situations later on. In the proof of the convergence of such sequences sometimes the following arguments turn out to be helpful. Let PI be an arbitrary metric on the set U (corresponding to the topology of U). The metric PI generates a Hausdorff metric on the set expc U. The last one induces a metric P2 on the space Cs{U) (by the embedding Gr). The metric P2 may be used to define a deflection a on the set of closed subsets of the space Cs(U): a{FI' F2 ) = inf{ {oo }U{e: e > 0, F2 ~ OeFd) (see §3.4, where we have considered the deflection on the set of compact subsets of a metric space). With this notation for Zo E Rc{U), K E expc U and c > 0 we have O{K, O«ZO)K, = {Z: Z E R(U), a{{ZO)K' ZK) < c}. By virtue of facts established in the step II of the proof of Lemma 7.1 we have: Assertion 7.1. A sequence {Zi: i = 1,2, ... } ~ R(U) converges to a space Zo E Rc(U) (in U or, that is equipotent, with respect to described topology of the space R(U)) if and only if a«ZO)K' (Zi)K) ---* 0 for every COmpactum K ~ U. • Now let
en
(7.2) Uo be an open subset of the set Uj
246
CHAPTER 7
(7.3) Ml ~ M2 ~ M3 ~ ... ~ U; (7.4) for every compactum K ~ Uo there exists an index i = 1,2, ... such that K ~ Mi. As examples of such sequences we may take Mi == Uo, or Mi == [Uo]. See also the proof of Theorem 3.2.2 (there X = Uo = U). Theorem 7.2. Let conditions (7.2 - 4) hold. Let Zo E Rc{U), {Zi : i = 1,2, ... } ~ R{U) and a{{ZO)Mi' (Zi)M) ~ O. Then the sequence {(Zd uo : i = 1,2, ... } converges to {Zo)uo in Uo. Proof. Take an arbitrary K ~ Uo and a sequence f3 = {Zi : i E B} ~ Cs(U) such that Zi E (Zi)K for every i E B. By (7.4) K ~ M i , beginning with some i = i o. By Lemma 2.3.4 for every i E B, i ~ i o, there exists a function z; E (Zo) Mi such that (7.5)
By the Corollary of Lemma 2.3.8 O,{K, co) ~ U for some co E (O, (0). We have a((ZO)Mi' (Zi)MJ + 2- i < co, beginning with some i = i 1 ~ i o· The last estimate and (7.5) imply that Gr(z;) ~ O(Gr(zi), co). Hence Gr(z;) ~ O,(K, co) E expc U. Since Z E Rc{U), the sequence {z;: i E B, i ~ id has a subsequence {z;: i E Bd converging to a function z* E Z. For i E Bl
Therefore the sequence {Zi: i E Bd converges to the function z*. This • gives what was required. The theorem is proved.
CHAPTER 8
PEANO, CARATHEODORY AND DAVY CONDITIONS
In Chapter 6 we have shown that in some quite simple cases the solution spaces of an equation satisfy the conditions which are suitable as axioms for the construction of a general theory. Although the theory is applicable to equations of very various types, for this time we have the possibility to use it only in quite narrow framework of the examples of Chapter 6. In this chapter we expand the sphere of applications of our theory. We expand it, in particular, to the cases which are covered by the classical theory of ordinary differential equations. In the using of the terms 'Caratheodory conditions' and 'Davy conditions' I follow a tradition of Russian mathematical texts. Many authors out of respect for the pioneering role of Caratheodory, reserve his name for denoting the most general situation here.
1. Kneser condition Let U be an open subset of the product ~ x ~n. The name of the following properties of a space Z in the title of the section:
~
C s (U) is mentioned
(k) for every point (t, y) of the set U there exists a number 6 > 0 such that for every point s E (t - 6, t + 6) the set {z(s): Z E Z, s, t E 7f(z), z(t) = y} is connected. The fulfilment of condition (u) implies the fulfilment of condition (k), because a one point set is connected. Lemma 1.1. Let Z E Rce(U). Let t E ~ and K be compact subsets of the set Un « -00, t] X ~n) (respectively, of the set Un ([t, 00) x ~n)). For every function Z E Z+ (respectively, Z E Z-) let t E 7f{z) ifGr(z)nK f:. 0. Then the set q,
= {z:
Z
E Z, SUP7l"(z)
= t,
(inf7l"(z),z(inf7f(z))) E K}
(respectively, q,
= {z:
Z
E Z, inf7l"(z)
= t,
(SUP7l"(z),z(SUP7l"(z))) E K})
248
CHAPTER 8
is compact. Proof. The lemma will be proved in main case. Assume the opposite. Then by Theorem 1.6.5 and remarks of §1.6 we can choose a sequence {CPk: k = 1,2, ... } ~ q, without convergent subsequences. By virtue of the compactness of the set K we can assume in addition that the sequence {(tk' CPk(t k )): k = 1,2, ... } ~ K, tk = inf7r(cpd, converges to a point (to, y). For every k = 1,2, ... fix an extension Zk E Z-+ of the function 'Pk (such an extension exists by Lemma 6.6.2). Apply Theorem 7.2.2 to the sequence of the functions {Zk: k = 1,2, ... } and to the sequence of the points {t k : k = 1,2, ... } with Zk = Z for every k = 1,2, .... Let z* E Z-+ be a function existing by Theorem 7.2.2. It takes the value y at the point to. Let {Zk: k E Ad be the corresponding subsequence. Since to E 7r(z*) and (to, z*(to» E K, t E 7r(z*). Take a number 8 > 0 such that [to - 8, t + 8] ~ 7r(z*). By virtue of the condition imposed in Theorem 7.2.2 on the subsequence {Zk: k E Ad we have [to - 8, t + 8] ~ 7r(Zk), beginning with some k = kl E AI, and on the segment [to - 8, t + 8] the sequence of the functions {Zk: k E AI, k ~ kl } converges uniformly to the function z*. We have tk E (to - 8, t + 8), beginning with some k = k2 ~ kl . Hence 7r( CPk) = [tk' t] ~ [to - 8, t + 8]. Theorem 3.5.4 implies the convergence of the sequence {CPk: k E AI, k ~ k2} to the function z*l[to,tj' This contradicts the choice of the sequence {CPk: k = 1,2, ... }. Thus our assumption is false and the lemma is proved. • Lemma 1.2. Under the hypotheses of Lemma 1.1 let Z E Rk(U) and the compactum K be connected. Then the set
S = {z(t):
Z
E Z-+, Gr(z) nK
i=
0}
is compact and connected (Le., it is a continuum). Proof. The lemma (as in the case of Lemma 1.1) will be proved in the main case. Theorem 3.5.4 implies the continuity of the mapping a : q, ---+ IRn , a(cp) = cp(t). By Lemma 1.1 the set q, is compact. Theorem 1.7.8 implies the compactness of the set S = a(q,). Assume that the set S is disconnected. Then S = Xl UX2, where the sets Xl and X 2 are nonempty, closed and disjoint. By virtue of the continuity of the mapping a and Theorem 1.7.5c the sets q,i = a-I(Xd, i = 1,2, are closed in q,. By Theorem 1.6.2 they are compact. Theorem 1. 7.8 implies the compactness of the sets Ki = U Gr(q,i), i = 1,2. By Theorem 1.6.1 these sets are ·closed. Theorem 1.3.4 implies the closedness of the sets K n Kl and KnK2. The nonemptiness of the sets KnK l and KnK2 is obvious in view of the nonemptiness of the sets q,l and q,2. If Z E Z-+ and Gr(z) n K i= 0, then 7r(z) 3 t. Thus the condition Z E Rce(U) and Lemma 6.6.2 imply that the sets K n Kl and K n K2 cover K. The connectedness of K implies the nonemptiness of the intersection (K n Kd n (K n K2)' So Kl n K2 i= 0.
Peano, Caratheodory and Davy conditions.
249
The projection of the nonempty set Kl n K2 ~ ~ x ~n in the first factor is compact. Its upper bound s « 00, see Remark 2.3.4) belongs to it. Take a point y E ~n such that (s, y) E Kl n K 2. We have s < t (the case s = t is impossible, because ({t} x ~n) n (Kl n K 2) = {t} x (Xl n X 2) = 0). Let for i = 1, 2 <1>;
= {z: Z E Z, 71'(z) = [s, t], z(s) = y, z(t) E Xd, K;* = UGr(<1>;).
The set <1>;, i = 1,2, is closed in the compactum {z: z E Z, sup 71'(z) = t, (inf7l'(z), (inf7l'(z» = (s,y)} (see Lemma 1.1). Therefore it is compact. By Theorem 1. 7.8 the set K; is compact. By the choice of s and by the definition of the sets K~ and K; we have K~ n K; = {( s, y)}. Therefore
for s' E (s, t). The sets ({ s'} So the set
({s'}
X ~n)
X ~n)
n (K: UK;)
n K~ and ({ s'} X ~n )nK; are nonempty.
= {z(s'): z E Z, s,s' E 71'(z), z{s) = y}
is disconnected, which contradicts condition (k). Thus our assumption is false. The set S is connected. The lemma is proved. Theorem 1.1. Let a sequence
•
{Zk: k = 1,2, ... }
~
Rcek(U)
converge in U to a space Z E Rc(U), Then Z E Rcek(U), Proof. I. By Theorem 7.2.1 Z E Re(U). It remains to prove that
Z E Rk(U), II. Let (to, Yo) be an arbitrary point ofthe set U. Fix numbers f.L > v > 0 according to Theorem 6.7.3 (for Z). Assume that for a number tl E [to - v, to + v] the set
S = {z(tl ) : z E Z, 71'(z) = [to - v, to
+ v],
z(t o) = Yo}
is disconnected: S = Xl U X 2 , where the sets Xl and X 2 are disjoint, nonempty and closed. Put for definiteness that to < t l . For i = 1,2 fix a function Zi E Z such that 71'(Zi) = [to, tIl, Zi(tO) = Yo and zi(td E Xi' Let H == Gr(zd U Gr(z2)' III. Let us show that, beginning with some k = kb we have: if z E Zk, 1r(z) == [a, b] ~ [to, to + v] and (a, z(a» E H, then Gr(z) ~ Op.Yo.
CHAPTER 8
250
Assume the opposite. Then there exist functions z;. E Zk, k E A, such that 7r(z;') = [ak, bk] ~ [to, to + v], Im(z;') ~ [OIL Yo] , (ak' z;'(ak)) E Hand Ilz;'(b k ) - Yoll = /-to By virtue of the convergence of the sequence {Zk : k = 1, 2, ... } to the space Z the sequence {z;': k E A} contains a subsequence converging to a function z E Z. Denote: 7r(z) = [a, b]. We have [a, b] ~ [to, to + v], z(a) = zi(a), where i = 1 or 2, and Ilz(b) - Yoll = /-to Consider the function z* E Z defined on the segment [to, b] by the formula *( ) = {Zi(t)
z t
z(t)
for t E [to, a], for t E [a, b].
The function z* cannot be extended to an element of Z-+ by virtue of our choice of /-to The obtained contradiction gives what was required. IV. Let WI and W 2 be disjoint neighborhoods of the sets Xl and X 2 , respectively, in the space ]Rn. Let
Bk
= {z: z
E
Zk, (inf7r(z),z(inf7r(z))) E H, SUP7r(z)
= td
for k = 1,2, .... By III
u Gr(Bd ~ (to - v, to + v) X GlLyo for k ~ k l . For k = kl' kl + 1, ... put X; = {z(tr): z E Bd. By Lemma 1.2 the set X; is connected. This observation and the nonemptiness of the sets Xl and X 2 imply the nonemptiness ofthe set X; \ (WI UW2 ). Fix a function z;. E Bk such that
By III
Gr(zZ) ~ [to - v, to
+ v]
x [OIlYO].
By virtue of the convergence of the sequence {Zk: k = 1,2, ... } to the space Z in U the sequence {z;': k = kl' kl + 1, ... } has a subsequence converging to a function z E Z. We have (a, z(a)) E H, where a = inf 7r(z), and z(tr) E ]Rn \ (WI U W 2 ), i.e., z(tr) rt. Xl U X 2 = S. Let z(a) = zi(a), where i = 1 or 2. Consider the function z* E Z such that 7r(z*) = [to, td and for t E [to, a] *( ) = {Zi(t) z t z(t) for t E [a, tl]. Let z** E Z-+ be an arbitrary extension of the function z*. By the choice of v we have [to-v, to+v] ~ 7r(z**). Next, z**(to) = Yo and z**(tr) = z(t 1 ) rt. S.
Peano, Caratheodory and Davy conditions.
251
This contradicts the definition of the set S. The obtained contradiction gives what was required. The theorem is proved. • Later on we will use Theorem 1.1 in proofs of the fulfilment of the Kneser condition. In order to follow this way we need to construct a space sequence (with properties which are mentioned in the statement of the theorem) converging to the space under consideration. In the scalar case the situation is simpler: Theorem 1.2. Let Uo be an open subset of the plane IR x IR and
Z E Rce(Uo). Then Z E Rcek(UO)' Proof. Take an arbitrary point (to, Yo) of the set Uo and find numbers J.I. > 11 > 0 according to Theorem 6.7.3. Our aim will be achieved when we show that for every point tl E (to - 11, to + 11) and for every points al < a2 of the set S = {z(td: Z E Z, ?fez) :3 to, t l , z(t o) = Yo} the segment [aI, a2] lies in S. Let for definiteness to < t l . Assume the opposite. Then there exists a point a E [aI, a2] \ S. Take an arbitrary function z E Z-+,
I
~a2 ~a
I I ,
to
S
Figure 8.1
tl
al
)
Na I
I
l )0
tl
Figure 8.2
the domain of which contains the point tl and which takes the value a at the point t l . For i = 1,2 fix functions Zi E Z such that ?f(Zi) = [to, tIl, Zi(t O) = Yo and zi(td = ai' Let I = [to, tIl n ?fez). The set I is either a segment or a half interval with an open left endpoint. If at a point s E I the graph of the function Z meets the graph of the function Zi, i = 1 or 2, then the function Z*(t) = {Zi(t) for t E [to, s] z(t) for t E Is, tIl
(1I'(Z*) = [to, tIl) belongs to Z. This contradicts the condition a ~ S (Figure 8.1). Thus Zl(t) < z(t) < Z2(t) for every point tEl (Figure 8.2). Hence Gr(ziI ~ [to - 11, to + 11] x [O~Yo]. By Lemma 6.6.1 the last fact may be true only when [to, tl] ~ ?fez). Then to E I and ZI(tO) < z(to) < Z2(t O) = Zl(t O ), that is impossible. The obtained contradiction gives what was required. The theorem is proved. •
252
CHAPTER 8
Theorem 1.3. Let Z E Rcek{U), (t, Yo) E U, t E [a, b]. Assume that if Z E Z-+, t E 7r{z) and z{t) = Yo then [a, b] ~ 7r{z). Then the set «P = {z : z E Z, 7r{z) = [a, b], z{t) = Yo} is connected. Proof. Let «Pa = {z/[a,tj: Z E «p} and «Pb = {Z/[t,bj: Z E «p}. The condition Z E R{U) implies, that «P = «Pa x «Pb (moreover this formula relates the topologies of these spaces too, see Theorems 3.5.4 and 3.5.5). By Lemma 1.1 the spaces «P, «Pa and «Pb are compact. The Corollary of Theorem 1.7.8 and Lemma 2.1.1 imply the upper semicontinuity of the mapping which is inverse to the projection of the product «Pa x «Pb into every factor. By Theorem 2.6.4 our assertion reduces to the cases t = a and t = b. Both these cases are analogous. For definiteness put t = a. Assume the opposite. Let «P = «P1 U «P2, where «P1 and «P2 are nonempty disjoint closed subsets of the compactum «P. Fix a countable dense subset T = {t k : k = 1,2, ... } of the segment [a, b] and for every k = 1,2, ... define the mapping h : «P -7 (1~n)k by the formula fk{Z) = (z{td,.·· ,z{tk )). Show that h{«P 1) n fk{«P 2) = 0 for some k = 1,2, .... If we assume the opposite, then for every k = 1,2, ... there are functions z~ E «Pi, i = 1,2, such that fk{Zl) = fk(Zn. By Theorem 1.7.8 the set U Gr{«p) is compact. It contains the graphs of all functions zL k = 1,2, ... , i = 1,2. Condition (c) implies a possibility to pass to subsequences {zl: k E A} and {z~ : k E A} converging, respectively, to functions Zl E «P1 and Z2 E «P2' For every k = 1,2, ... we have Zl{t k ) = Z2{t k ). The density of the set T in the segment [a, b] and Lemma 1.7.4 imply the coincidence of the functions Zl and Z2. Therefore «P1 n «P2 =J 0, which contradicts the choice of the sets «P1 and «P2' Thus for a number k the set fk{«P) is disconnected. To continue the reasoning it is convenient to word the obtained result in an other way: there exist such points Sl < ... < Sk of the segment [a, b] that for the mapping f : «P -7 (1~n)k, f{z) = (z{sd, ... , z{sd), the set f{«p) is disconnected. Fix such a mapping f, for which the number k is the smallest of all possible ones. Lemma 1.2 immediately implies, that k ~ 2. Let the mapping 9 : «P -7 (Rn)k-1 be defined by the formula g{z) = (z{sd, ... , z{sk-d)· Let h : f (<
•
Peano, Caratheodory and Davy conditions.
253
Consider the vectors el = {1,0}, e2 = {O, -I}, e3 = -el, e4 = -e2 and the subsets Ml = {(x,y): x ~ 0, -x:::;; y:::;; x}, M2 = ((x,y): y ~ 0, -y:::;;x:::;;y},M3 ={(x,y): x:::;; 0, x:::;;y:::;;-X},M4={(X,y): y:::;;O, y:::;; x:::;; -V} of the plane. Let U = jR X jR2. Let Zi be the solution space of the equation y' = ei, i = 1,2,3,4. Define the space Z by the condition: ZIRxMi = ZilRXMi for -' t i = 1,2,3,4 (see Figure 8.3 and §9.5) below. Evidently such a space Z E R(U) exists and is defined uniquely by this condition. The reader can show that Z E Rce(U) and that at the point (t, 0, 0) the Kneser condition is not fulfilled. Notice now that Theorem Figure 8.3 7.2.3 implies Lemma 1.3. Let Z E Rce(U) and'Y be an open cover of the set U, for every element V of which Zv E Rk(V), Then Z E Rk(U), •
,,
..,
, ", -,
-,
"
2. Caratheodory and Davy conditions In this section we investigate a class of equations and inclusions studied by Caratheodory (case of equations, see [CLD and by Davy (a larger case of inclusions, see [DaD. Start with preliminary remarks. See also [LR] in the connection with the approximation of right hand sides. Lemma 2.1. Let multi-valued mappings Fk, k = 0,1,2, ... , be defined on a topological space X, take their values in the Euclidean space Rn and be upper semicontinuous. Assume that for every x E X the set F(x) = CC(U{Fk(X): k = 0,1,2, ... }) is compact and
(2.1) for every neighborhood OFo(x) of the set Fo(x) in ~n there are a neighborhood Ox of the point x in X and a number ko = 1,2, ... such that U{Fk(OX): k = ko, ko + 1, ko + 2, ... } ~ OFo(x). Then the mapping F : X ~ jRn is upper semicontinuous. Proof. Take an arbitrary point x E X and an arbitrary neighborhood W ofthe set F(x). By Corollary of Lemma 2.3.8 026F(X) ~ W for a number 6> O. For the neighborhood 06FO(X) of the set Fo{x) find a neighborhood Ox and a number ko according to (2.1). For k = 0,1, ... , ko -1 by virtue of the upper semicontinuity of the mapping Fk there exists a neighborhood Vk of the point x in X such that Fk{Vk ) ~ 06Fk{X). For t E oxnv1n·· ·nVko - l
254
CHAPTER 8
we have
By Lemma 2.4.3 the set OcF(x) is convex. Lemma 2.4.5 implies the convexity of the set [OcF(x)]. The definition of the closed convex hull implies the inclusion F(t) ~ [OcF(x)] ~ W. The lemma is proved. • Let as before U denote an open subset of the product ~ x ~n. Lemma 2.2. Let Fk : U --t ~n, k = 1,2, ... , be multi-valued mappings. Let Fl(X) 2 F2(x) 2 F3(X) 2 ... and F(x) = n{Fk(X): k = 1,2, ... } for every point x E U. Then D(F) = n{D(Fk): k = 1,2, ... }. Proof. Evidently D(F) ~ D(Fk ) for every k = 1,2,3, .... Therefore D(F) ~ n{D(Fk): k = 1,2, ... }. Prove the inclusion n{D(Fk): k = 1,2, ... } ~ D(F). Let Z E n{D(Fk): k = 1,2, ... }. If 1l'(z) consists of one point only then Z E D(F) by definitions. Let 1l'(z) contain more than one point. For every k = 1,2, ... we can point in the segment 1l'(z) a subset Mk of measure zero such that for every point t E 1l'(z) \ Mk the derivative z'(t) exists and belongs to the set Fk(t, z(t». By Theorem 4.1.2 the measure of the set M = U{Mk: k = 1,2, ... } is equal to zero. For every point t E 1l'(z) \ M we have z'(t) E n{Fk(t, z(t»: k = 1,2, ... } = F(t, z(t». Thus z E D(F). The lemma is proved. • Lemma 2.3. Let a = {Zk: k = 1,2, ... } ~ Rc(U), Zl 2 Z2 2 ... and Z = na. Then the sequence a converges in U to the space Z. Proof. The fulfilment of the condition a E s(U) follows from Theorem 3.6.1, because all elements of a lie in Zl E Rc(U). By Lemma 6.4.1 and Theorem 1.6.1 elements of the sequence a are closed in Cs(U). Referring to remarks of Example 1.5.3 and to Lemma 7.1.4 completes the proof. • The following assertion is obvious. Lemma 2.4. If a sequence {Zk: k = 1,2, ... } of subspaces of the space Cs(U) converges in U to a space Z ~ Cs(U) and Z; ~ Zk for every k = 1,2, ... then the sequence {Z;: k = 1,2, ... } converges to the space Z in U. • Let us now turn to the main topic of this section. Let (beginning here and to the end of the section) cp denote a non-negative function which is defined and locally Lebesgue integrable on an interval (a, b), -00 :::;; a < b :::;; 00, of the real line and U ~ (a, b) x ~n. Denote by Qd('P)(U) the set of all multi-valued mappings F : U --t ~n with compact convex values satisfying the conditions: a) there exists a single valued mapping f : U --t ~n such that for every y E ~n the mapping fY is measurable and f(t, y) E F(t, y) ~ 0,(0, cp(t» for every (t, y) E Uj
Peano, Caratheodory and Davy conditions.
255
b) for every t E (a, b) the mapping Ft is upper semicontinuous. These are the Davy conditions. For F(t,y) == {j(t,y)} they go into Caratheodory conditions. The function cp is called a (Davy, respectively, Caratheodory) majorant of the right hand side F. Notice that Qd('P)(U) ~ Q*(U) (see §6.4 and Theorem 2.5.2). By results of §§6.4 and 6.5 for F E Qd('P)(U) the solution space D(F) satisfies conditions (c) and (n). Now our goal is to prove the fulfilment of conditions (e) and (k). For t E (a, b) and Y E ~n put if (t, y) E U, in the opposite case. Evidently for every y E ~n the function all t E (a, b) we have IIf~(t)1I :::;; cp(t). Define on the space ~n the function:
Jr
.
If
is measurable. For almost
n
Ilx - yll : :; k'
in the opposite case, where k = 1,2, ... , Y E ~n. By Theorem 1.7.6 it is continuous. A simple direct calculation with an analogous using of Theorem 1.7.6 allows to check the existence and the continuity of its derivatives. Let ~(x,k) = (~, ... ,~) for x= (k1, ... ,kn ) E Nn and k = 1,2, .... ~or every k = 1,2,... and y = (Yl, ... , Yn) E ~n there exists x = (k 1, ... , k n ) E Nn such that IkYi - kil < 1 for i = 1, ... ,n. Then
IIY -
~(x, k)1I
1/
2
2
...;n < k' n
= k Y (kYl - kd + ... + (kYn - kn) :::;; k G:k,{(x,k) (y) > o.
On the other hand, if O(y, n/k) n O(~(x, k), n/k) 1= 0, where x = (k 1, ... , kn ) E Nn, k = 1,2, ... , then p(y,~(x, k)) < 2n/k,
IkYl - kl l2 + ... + IkYn -
knl2 < 4n 2 , IkYl -
kll < 2n, ... , IkYn
- knl < 2n.
Only a finite number of x E Nn satisfy the last condition. Therefore the function 13Z(x) = E{G:k,{(x,k) (x) : x EN} is continuous and has continuous derivatives. For k = 1,2, ... , x E Nn and x E ~n put (.l
fJk,x
()
x
=
G:k,{(x,k) (x)
13Z(x)'
CHAPTER 8
256
These functions are continuous. They have continuous derivatives. For every point x E IRn:
(2.2)
!3k,,.,(X)
~
0;
(2.3) there exists a neighborhood Ox of the point x, on which only finite number of functions of the family {!3k,,.,: x E Nn} do not equal identically zero;
For every k = 1,2, ... define on the set (a, b) x IRn the function:
Let now (to, Yo) E U. Take an arbitrary number "I > 0 satisfying the condition [to - "I, to + "II x O,(Yo, "I) ~ u. By virtue of the compactness of the set O,(Yo, "I) and (2.3) the set A = {x: x E Nn, SUP!3k,,.,(O,(yo, "I)) > O} contains only a finite number lk of elements. Put mk = sup{ 8:~,>< (x) : x E A, j = 1, ... ,n, x E O,(Yo, "I)} J « 00, see Remark 2.3.4). For every t E [to - "I, to + "II the norms of all derivatives of the functions ik(t, y) in yare not greater than lkmkcp(t). By Lemma 6.5.4 (2.5) for every point p, q E OT/Yo, By virtue of our assumptions, of Theorems 4.2.6 and of (2.5) the function h satisfies on the set Uo = (to -"I, to +1J[ xOT/Yo conditions a) and b) of §6.5. Remarks in Example 6.5.1 imply the membership D(fk, Uo) E Rceu(Uo). By Lemmas 6.5.2 and 6.5.3 and the arbitrariness in the choice of the point (to, Yo) we have D(h, U) E Reu(U). This fact and remarks in the beginning of the reasoning imply the membership D(A, U) E Rceu(U), Show that for every t E (a, b) at every point x E Ut the sequence of the mappings Fo = Ft , Fk = (fklu)t for k = 1,2, ... satisfies (2.1). Let o Fo (x) be an arbitrary neighborhood of the set Fo (x) in the space IRn. By Corollary of Lemma 2.3.8 OeFO(x) ~ OFo(x) for some c > O. By virtue of the upper semicontinuity of the mapping Fo there exists a neighborhood Ox of the point x in Ut such that Fo(Ox) ~ OeFO(x). There are numbers 8 > 0 such that 025X ~ Ox and ko = 1,2, ... such that ::a < 8. If Y E Oe x, k = k o, ko + 1, ko + 2, ... , x E A and !3k,x(y) =f= 0, then O(~(x, k), V :3 y. By virtue of the triangle inequality ~(x, k) E 025X ~ Ox ~ UtI Hence (2.6)
J.(t, ~(x, k» = J(t, ~(x, k))
E
OeFo(x).
Peano, Caratheodory and Davy conditions.
257
By (2.2) and (2.4) the point Ik{t, y) belongs to the convex hull of the set {!(t,~(x, k)): x E Nn, f3k,x(Y) -=I O}. By Lemma 2.4.3 the set OeFO(x) is convex. Now (2.6) implies that Ik(t, y) E OeFO{x) ~ OFo(x). In view of the arbitrariness of k = ko, ko + 1, ko + 2, ... this means the fulfilment of (2.1). Lemma 2.1 implies the upper semicontinuity in y of the mappings
Gk : U
-+ ~n,
Gk(t,y)
= cc(F(t,Y)U{!i{t,y): i = k,k+1,k+2, ... }), k =
1,2, .... Thus the mapping G k satisfies condition b). Evidently the mapping satisfies condition a) too. Therefore G k E Qd(rp)(U), D{G k) E Rc{U). Consider an arbitrary point (t, y) of the set U. We noticed above that the sequence of the mappings Fo = Ft , Fk = (fklu)t, k = 1,2, ... , satisfies (2.1). By Lemmas 2.4.2 and 2.4.3 we have: F(t,y) = n{OeF{t,y) : c > O} 2 n{ Gk(t, y): k = 1,2, ... }. Since the inverse inclusion is obvious, F(t, y) = n{Gk(t, y): k = 1,2, ... }. By Lemma 2.2 D(F) = n{D{G k ) : k = 1,2, ... }. By Lemmas 2.3 and 2.4 we obtain the convergence in U of the sequence {D(fk, U): k = 1,2, ... } to the space D(F). The proved convergence D(fk, U) -+ D(F) (in U), our remarks, and Theorem 1.1 imply: Theorem 2.1. Let FE Qd(rp)U. Then D(F) E Rcekn(U), Now give several simple examples in order to familiarize ourselves with the scheme of the using of Theorem 2.1. Example 2.1. The right hand side of the scalar equation I
y=
sin(t + y)
vm
is defined on the entire plane except the line t = O. The manner in which
we define it on this line does not affect on the notion of a solution of this equation. All changes are possible only for the unique value of the argument t = O. A solution need satisfy our equation almost everywhere only (see §6.2). Under such a change a solution remains a solution and new solutions do not appear. For instance, we can extend the right hand side of the equation to a function I(t, y) by putting 1(0, y) = 0 for every y E R. The function 1 satisfies the Caratheodory conditions. As a majorant 'P here we can take the function for t -=I 0 for t = O. The function 'PI is locally Lebesgue integrable, because almost everywhere (more precisely except the point t = 0) it is the derivative of the non-
258
CHAPTER 8
decreasing function w(t) = 21tl t sgn t, see Lemma 4.9.5. Recall that -I
sgnt
=
{
~
if t < 0, if t = 0, if t > 0.
By Theorem 2.1 the solution space belongs to Rcekn(lR x 1R). Notice that all solutions of this equation are absolutely continuous. This is a direct consequence of Lemma 4.7.6. Example 2.2. Consider now the differential inclusion with the right hand side F(t ) = [sin2 (t + y) sin2 (t + y)].
JItT
2J1tT'
,y
As in the previous example the right hand side is not defined for t = 0. Define it for t = by putting F(t, y) = {o}. We obtain the differential inclusion with the same solution set, see the analogous reasoning in Example 2.1. The function
°
for t
# 0,
for t = 0, as the function f of Davy conditions. Thus the solution space of the differential inclusion y' E F(t, y), or (with the other notation) the solution space of the inequality
belongs to Rcekn(IR x 1R). As in the previous example all solutions of our inclusion (respectively, inequality) are absolutely continuous. Example 2.3. Give an example of a simplest situation which cannot be investigated completely by the application of Theorem 2.1. The space of Example 1.1 is the solution space of the differential equation with the right hand side
{ {l,O} f(t,x,y)=
{-1,0} {O,-I}
{O,I}
for for for for
x ~ 0, x < 0, y > 0, y < 0,
-x ~ y ~ x, x ~ y ~ -x, -y < x < -y, y < x < -yo
Peano, Caratheodory and Davy conditions.
259
The equation u' = f(t,u) (here u = (x,y» satisfies all Caratheodory conditions, except the condition of the continuity of the right hand side in u. This difference with hypotheses of the theorem leads to the non-fulfilment of the Kneser condition. Pass now to the continuity of the dependence of solutions on parameters of the right hand side. Since the question about the uniqueness of solutions we leave apart, in view of Theorem 2.5.1 it is sufficient to study the question about the convergence of sequences of solution spaces. In fact, the largest part of this investigation is done in §7.1 and it remains to compare the obtained results with the Caratheodory and Davy conditions. Lemma 2.5. Let F : U - t IR n be a multi-valued mapping with compact convex values. Let F(t, y) ~ 0,(0, cp(t» for every point (t, y) E U and the mapping Ft be upper semicontinuous. Then the mapping F is weakly continuous with respect to the space Cs(U). Proof. Let [a, bJ be a segment of the real line. Let functions z, Zk E C.(U), k = 1,2, ... , be defined on the segment [a, bJ and Zk - t z. Let elements of a sequence {ak: k = 1,2, ... } be Denjoy integrable functions defined on the segment [a, bJ and satisfy for almost all t E [a, bJ the condition D:k(t) E F(t, Zk, (t». Let the sequence of the functions
{ A.(t)
~
j
a.(s)ds: k
~ 1,2 ... }
converge uniformly to a function A. All functions A k, k = 1,2, are solutions of the inequality Ily'(t)1I ~ cp(t). By Lemma 6.3.2 the function A solves this inequality too. By our definitions and Lemma 6.3.3 for every point t from a subset M of full measure of the segment [a, bJ we have ak(t) = A~(t) E F(t, Zk(t» for of all k = 1,2, ... , the derivative A'(t) exists and (2.7)
A'(t)En{cc{A~(t):
k=j,j+1, ... }: j=1,2, ... }= =n{Cc{ak(t): k=j,j+1, ... }: j=1,2}.
For every neighborhood V of the point z{t) we have Zk{t) E V, beginning with some k. Theorem 2.5.2 and (2.7) imply the belonging A'(t) E cc(Ft(V». By virtue of the upper semicontinuity of the mapping F t and Theorem 2.5.1 the conditions ak(t) E F{t, Zk{t» and Zk{t) - t z{t) imply the nonemptiness ofthe set F{t, z{t». Remarks of §6.4 imply now the inclusion F(t,z{t» ;2 n{Ft{V): V is a neighborhood of the point z{t) in Ud 3 A'{t). The lemma is proved. •
CHAPTER 8
260
Lemma 2.6. Let hypotheses of Lemma 2.5 hold. Let a sequence of multivalued mappings Fi : U --t IRn, i = 1,2, ... , converge in U with respect to a space Z ~ Gs(U) to a mapping F. Then {D(Fi) n Z: i = 1,2, ... } E s(U). Proof. Let K be an arbitrary compact subset of the set U. Let Zi E D(Fi' K) n Z for i = 1,2, ... and {'Pi: i = 1,2, ... } be a sequence of Lebesgue integrable positive functions such that
Fi(t, Zk(t)) ~ O(F(t, Zi(t)), 'Pi(t)) and
J
< bZ,K(F,Fi ) +
'Pi(s)ds
t·
7r(Zi)
If now 0:, {3 E 1f(Zi), i = 1,2, ... , and 0: ~ {3, then (3
II Z i({3) - zi(o:)11
JIlz~(s)llds ~J +J ~J + ~
a
(3
(2.8)
(3
'P(s)ds
'PI (s)ds
(3
bZ,K(F,Fi ) + Iii.
'P(s)ds
a
For arbitrary c > 0 take an indexj = 1,2, ... such that bZ,K(F, Fi )+l/i < ~ for i = i + 1, i + 2, ... (it may be done by virtue of the convergence of the sequence {Fi: i = 1,2, ... } to F in U with respect to Z) and a number b > 0 such that
J
'P(s)ds <
M
~,
J
'Pi(s)ds
<
~
M
for i = 1, ... ,j and for every segment M ~ 1f(Zi), i = 1, ... ,j, of the length < b (it may be done by virtue of the absolute continuity of the Lebesgue integral). If now 0:, {3 E 1f(z;), i = 1,2 ... , and 10: - {31 < b, then Ilz({3) - z(o:)11 ~ c by (2.8). It means the equicontinuity of the sequence {Zi : i = 1,2, ... }. Hence {D(Fi) n Z : i = 1,2, ... } E s(U). The lemma is proved. • Lemmas 2.5 and 2.6 and Theorem 7.1.2 imply immediately: Theorem 2.2. Let F E Qd(cp)(U) n Q/(U). Let a sequence of multivalued mappings Fk : U --t IRn, k = 1,2, ... , converge with respect to the space Gs(U) to the mapping F. Then the sequence {D(Fk): k = 1,2, ... } converges in U to the space D(F). •
Peano, Caratheodory and Davy conditions.
261
By Theorem 4.13.3 every single valued mapping cP E Qd(cp)(U) belongs to the set Q/(U). So: Theorem 2.3. Let a single valued function f : U ~ ]Rn belong to Qd(CP)(U). Let a sequence of multi-valued mappings Fk : U ~ IRn, k = 1,2, ... , converge in U with respect to the space Cs(U) to the function f. Then the sequence {D(Fk): k = 1,2, ... } converges in U to the space D(f). • Remark 2.1. To prove the convergence of a sequence of multi-valued mappings Fk : U ~ IRn, k = 1,2, ... , to a mapping F : U ~ Rn we have a simple sufficient condition. Let positive Lebesgue integrable functions CPk, k = 1,2, ... , be defined on a segment [a, b]' U ~ (a, b) x Rn, Fk(t, y) ~ Of(F(t, y), IPk(t») for every (t, y) E U and k = 1,2, .... Let IPk(s)ds ~ 0 as k ~ 00. Then the sequence {Fk: k = 1,2 ... } converges in U with respect to C s (U) to the mapping F. Proof of this fact is obvious, see the definitions in §7.1. Many assertions about the continuity of the dependence of solutions on parameters may be obtained as partiCular case of these theorems. Give an example of their specific use. Example 2.4. The equation
J:
, y=
sin(t + y)
Jltf
is considered in Example 2.1. In an analogous way we may investigate the equation , _ sin(t + y) I la yJltf +t, where -1 < Q < 1. Theorem 2.3 implies the continuity of the dependence of solutions on the parameter Q. Use Theorem 2.3 and consider the function 1/JaIJ(t) = Iiti a - ItlIJI to estimate the deflection of the right hand side (see the function CPk in Remark 2.1). We obtain the needed estimate from the Convergence 1
f
-1
f Isa 1
1/JaIJ(s)ds
=2
0
sIJlds
=
21_1_ -_1_1 ~ Q+1 .8+1
0
as .8 ~ Q. The same calculations allow to obtain an analogous result for the equation ,_ sin(t+y) Ilah( ) y Jltf + t t, Y , where h( t, y) is an arbitrary continuous bounded function.
262
CHAPTER 8
Se also Theorem XII.2.3 below.
3. Localization principle The investigation of properties of solution spaces may be essentially simplified by a possibility to do it locally with a further carryover of obtained results to the entire domain of the right hand side. A possibility of such a carryover is based on Lemmas 6.6.2 (for (e», 6.6.2 (for (u», 1.3 (for (k) under the fulfilment of (c) and (e)), on Theorem 7.6.4 (for the convergence of space sequences) and on its Corollary (for (c». For instance, this scheme of reasoning was used in the proof of Theorem 2.1. Point a corresponding result for condition (n). We have in mind its other applications too. The following its a little more general statement will be convenient. Let U be an open subset of the product IR x IRn. Lemma 3.1. Let a space Z E R(U) consist of generalized absolutely continuous functions. Let 'Y be a family of open subsets of the set U, for every element V of which Zv E Rn(V). Let J.Lz(U \ (U'Y» = O. Then Z E Rn(U). Proof. Let Zk E Z, 7r(Zk) = [a,b], k = 1,2 .... Let the sequence {Zk: k = 1, 2 ... } converge uniformly to a function Z E Z. Denote by v the family of all segments lying in [a, b] and satisfying the condition Gr(zlr) ~ V for some V E 'Y.
The hypotheses of the lemma imply that the measure of the set Mo = [a, b] \ (U{ (I) : I E v}) is equal to zero. By Theorem 1.6.10 the family v contains an (at most) countable subfamily Vo = {II, 12, ... } with U{(I): I E v} = U{(I): IE vol. For every j = 1,2, ... we can associate a subset Mi of measure zero of the segment I j such that for every point t E Ii \ M j all derivatives z'(t), z~(t), k = 1,2, exist and satisfy condition (6.3.1). By Theorem 4.1.2 the measure of the set M = U{ Mi : j = 0, 1, 2, ... } is equal to zero. At every point t E [a, b] \ M condition (6.3.1). holds. The lemma is proved. • In framework of the purpose of this section notice: Corollary. Let Z E R(U). Let 'Y be an open cover of the set U, for every element V of which Zv E Rn(V). Then Z E Rn(U). (The estimate J.L(U \ (U'Y)) = 0 is obvious. The generalized absolute continuity of elements of Z follows easily from Lemmas 4.9.1 and 7.5.1: in the last one for Z E Z we need take as 'Y the family of all connected open subsets of the segment 7r(z), on which the function z is generalized • absolutely continuous.) Look now what the idea of the localization of the investigation gives in particular cases.
Peano, Caratheodory and Davy conditions.
263
Example 3.1. Let U ~ (a, b) x Rn, where -00 ~ a < b ~ 00. Let functions aj, j = 1, ... , d, be locally Lebesgue integrable on the interval (a, b). Let functions h : U --t ~n, j = 1, ... ,d, be continuous. Let
Consider the equation
By remarks of §2.1 for every t E ~ the function It is continuous. By Theorem 4.2.6 for every Y E ~ the function IY is measurable. We can use Theorem 2.1 if we present an integrable majorant of the right hand side. Under our assumptions this turns out to be impossible. However Theorem 2.1 is applicable locally. Our general remarks about the localization principle allow to complete the investigation. In fact every point x of the set U possesses a neighborhood V with the compact closure lying in U (see Lemma 2.4.2). By Remark 2.3.4 for every j = 1, ... , d and for mj = sup{lI/j(t, Y)II: (t, y) E [V]} we have mj < 00. By Theorem 4.7.1 and remarks of Example 4.7.7 the function cp(t) = mI!al(t)1 + ... + mdlad(t)1 (defined on (a,b)) is locally Lebesgue integrable. It majorizes the function I on the compactum [V] and all the more on the set V: 11/(t, y)1I ~ cp(t) for (t, y) E V. Naturally, remarks about the function I remain true for the function 9 = I Iv: for every t E ~ the function gt is continuous and for every y E ~n the function gY is measurable. The function 9 has a locally Lebesgue integrable majorant. Thus the function 9 satisfies the Caratheodory conditions. By Theorem 2.1 the space D(J, V) = D(g) belongs to Rcekn(V), Such a neighborhood V may be pointed for every x E U. Therefore D(J) E Rcekn(U), Example 3.2. Let under assumptions of Example 3.1 the functions Ii, j = 1, ... , d, have continuous partial derivatives in y. For arbitrary (to, Yo) E U find c > 0 such that [to - c, to + c] x [OEYO] ~ U. Let K
=
[to - c, to
+ c]
x [OEYO] and M
= SUP{II~(t,Y)II:
(t,y) E K,
n}.
By Remark 2.3.4 M < 00. The function CPl(t) = Mn(lal(t)1 + ... + lad(t)l) is locally Lebesgue integrable (see Theorem 4.7.1 and Example 4.7.7). By Lemma 6.5.4
j == 1, ... , d, k = 1, ... ,
II/(t,p) - f(t,q)1I ~ Mn(l a l(t)1
+ ... + lad(t)I)lIp - qll =
cpdt) lip -
qll
for t E (to - c, to + c) and p, q E OeYo. Thus on the set V = (to - c, to + c) X OEYO the right hand side of the equation y' = I(t, y) satisfies condition b) of §6.5. Our new set V satisfies
264
CHAPTER 8
all conditions imposed on the set V in Example 3.1. Thus on the set V the right hand side of the equation under consideration satisfies all hypotheses of Theorem 6.5.1. So D(f, V) E Rceun(V). We can point such a neighborhood V for every point of the set U. Remarks and lemmas of this section imply that D(f) E Rceun(U). Example 3.3. Let real functions a and (3 be defined and be Lebesgue integrable on the interval (a, b), -00 ~ a < b ~ 00. Evidently the right hand side of the equation y' = a(t)y + (3(t) satisfies all condition imposed in Example 3.2 (with d = 2, f1(t,y) == y and f2(t,y) == 1). Thus the solution space of this equation belongs to Rceun(V), where V = (a, b) x ]Rn. Let to E (a, b) and A(t) = It: a(s )ds. A direct verification can show that the formula (3.1)
y(t) =
1
(YO +
e- A(')/3(S)dS) eA('1
defines a solution of our equation with the initial value y(t o) = Yo. By virtue of condition (u) this is the unique solution with this initial value. Notice a particularity of the formula (3.1): it defines the function y(t) on the entire interval (a, b). Let now z be an arbitrary solution of our equation and t1 E 71'(z). The function Zl (t)
= ( Z (t) 1 +
f.' a(u)du j e-.r a(u)du(3( s)d) se t
'I
'I
tt
is defined on the interval (a, b) and is also a solution of the equation under . consideration with the initial value zl(td = z(td. Therefore by virtue of condition (u) it coincides with z on 71'( z). Since to E 71'( Zl) (and Zl is a solution), for some value Yo the function Zl is defined by (3.1). Thus the formula (3.1) represents all solutions of our equation. Expressions of the type
remark.
Peano, Caratheodory and Davy conditions.
265
Let a non-negative function 0, e E IRn and G(t) = {u: u E IRn , (u -
= 0/(0,
for t E (a, b). Show that every solution z of the differential inclusion z' (t) E H (t) is absolutely continuous. Let 7r(z) = [e, dj and M be a set of all points t E [e, dj, at which the derivative z'(t) exists and Ilz'(t) -
Lebesgue integrable. By Lemma 4.9.2 the function t
z*(t) = z(t) -
J
'l/J1(s)ds
c
is generalized absolutely continuous on [e, dj and
(z*)'(t) = z'(t) - z'{t)
+
(almost always) for t E M and (z*)'(t) = z'(t) E H(t) for t rt M. Since (z*)'(t) rt 0/(
~
.!.'l/J2(t) a
Lemma 4.7.6 implies the Lebesgue integrability of the function (z*)'(t) -
J t
z.... (t)
= z"(t) -
c
CHAPTER 8
266
is absolutely continuous. By Theorem 4.7.1 this implies the absolute continuity of the function z* (t) and the absolute continuity of the function t
z(t) = z*(t)
+
J
'1Pt (s)ds,
c
which gives what was required. Evidently Remark 4.1. If U is an open subset of the product ~ x ~n, Z E R(U) and, is an open cover of the set U, for every element V of which the set Zv consists of absolutely continuous functions, then the set Z consists of absolutely continuous functions too. 5. Continuously differentiable solutions. Peano and Picard conditions In this section we speak mainly about solutions of a differential equation y' = f(t, y) (with the right hand side f defined on an open subset U of the product ~ x ~n). Lemma 5.1. Let z E D(f), 7l'(z) = [a, b], where a < b. Let a function
fIGr(z) be continuous. Then for every t E [a, b] the derivative z'(t) exists and is equal to f(t, z(t)). We speak here not about the approximate derivative. We speak here about the derivative in the common sense and as a consequence we obtain the continuity of the derivative. Proof. Let M be a subset of full measure of the segment [a, b]. Let the approximate derivative z'(s) be defined and satisfy the equality z'(s) == f(s, z(s)) for every s E M. Take an arbitrary point to of the segment [a, b]. By Theorem 4.10.1
z(q) - z(P) E cc(cp(M n [p, q])) q-p
for every p < q, a ~ p ~ to ~ q ~ b. Pass here to the limit as p, q - t. Lemma 2.5.2 and the continuity of the function fIGr(z) imply the existence of the limit of the quotient in the left-hand side (i.e., the existence of the usual derivative z'(t)) and its equality to f(t, z(t)). The lemma is proved .• The proved lemma implies immediately that under the assumption of the continuity of the right hand side (this is just the Peano condition) every solution of our differential equation is continuously differentiable. By remarks of Example 3.1 in this case the solution space satisfies conditions
Peano, Caratheodory and Davy conditions.
267
(c), (e), (k) and (n). Under these assumptions the last condition is fulfilled in the following reinforced version: (n*) all functions from (the solution space of the equation under consideration) Z are continuously differentiable and if Zk E Z, 7r(Zk) = I, k = 1,2, ... , and the sequence {Zk: k = 1,2, ... } converges uniformly to a function Z E Z, then the sequence {z~: k = 1,2, ... } converges uniformly to the function z'.
(As in (n*) we speak about the convergence ofthe sequence offunctions I(t, Zk(t)), then we can deduce (n*) from Theorems 3.3.1 and 3.5.3.) The Picard condition consists in the requirement of the continuity of the right hand side I of the equation under consideration and of the existence of a number M ;;:: 0 such that II/(t, Yl) - I(t, Y2)11 ~ MllYl - Y211 for every two points (t, Yl) and (t, Y2) (with coinciding first coordinates) of the set U. Properties of equations satisfying the Peano condition, naturally, remain true for equations satisfying the Picard condition. In addition the second part of condition b of §5.5 is fulfilled too. Therefore the solution space of the equation with the Picard condition satisfies condition (u). This remain true for equations with right hand sides locally satisfying the Picard condition (see §3). In fact the requirement of the continuity of the right hand side for the using of Lemma 5.1 is too strong. We may point out other cases of its application. Let x = (t, y) E U and Z ~ Cs(U). The set G ~ U is called the left Z -neighborhood (respectively, right Z -neighborhood, Z -neighborhood) of the point x, if it contains the point x, the set G \ {x} is open with respect to the Euclidean topology and for every function Z E Z, Gr(z) 3 x, there exists a number {j > 0 such that Gr(zl ... (z)n[t_6,t)) ~ G (respectively, Gr(zl ... (z)n[t,t+6)) ~ G, Gr(zl7r(z)n[t-6,t+6)) ~ G)). The set GEx (where c > 0) is a Z-neighborhood of the point x for every Z ~ Cs(U). Every Z-neighborhood of the point x is its left and right Z-neighborhood. If G 1 is a left and G 2 is a right Z-neighborhoods of the point x, then the set G 1 U G 2 is a Z-neighborhood of the point x. The set of all Z-neighborhoods of the point x is directed to the point :t (see §1.2). The topology rZ generated by the system of neighborhoods x E U} coincides on the graph of every function from Z with the Euclidean topology. Therefore Lemma 5.1 implies Theorem 5.1. Let a (single valued) function f : U --+ ]Rn be continuous with respect to the topology rD(f). Then for every function Z E Z (with the domain containing more than one point) at every point t E 7r(z) the (Usual) derivative z'(t) exists, is equal to f(t, z(t» and the function z'(t) is COntinuous. • z~(t) =
r;
{r;:
268
CHAPTER 8
The direct using of Theorem 5.1 needs a full description of the topology In order to obtain the description in most cases we need know well properties of solutions. Already the situation may be easily corrected by the following remark. TD(f).
Let a topology T be a part of the topology TD(f) (i.e., T ~ TD(f)). If a function is continuous with respect to the topology T then it is continuous with respect to the topology TD(f) too. Therefore to apply Theorem 5.1 it is sufficient to point a topology or a system of neighborhoods of points of U, which estimates properties of the right hand side of the equation in a needed way. A left contingent cont,(Z,x) (respectively, right contingent contr(Z,x), contingent cont(Z, x)) of the space Z ~ Cs(U) at a point x = (t, y) E U is the set
.
hm top sup 0--+0
{z(s) - z(t) S -
t
: z E Z, s, t E 7l'(z),
S
E (t - e, t), z(t) = Y
}
(respectively, . {z(s) - z(t) : zEZ, s,tE7l'(Z), SE(t,t+c), z(t)=y } , 11m top sup 0--+0 s- t
.
hm top sup 0--+0
{z(s) - z(t) S -
t
s E (t - c, t) U (t, t
: z E Z, s,t E 7l'(z),
+ c),
z(t) = y} (= cont,(Z, x) U contr(Z, x))).
Let F : U -+ ]R.n be an arbitrary multi-valued mapping, G be a left D(F)-neighborhood of a point x E U. Theorem 4.10.1 implies that cont,(D(F), x) ~ cc(F(G)). By Theorems 2.4.1 and 1.7.2 we obtain: Theorem 5.2. Let F : U -+ ]R.n be a multi-valued mapping. Let for
a left D(F)-neighborhood G of a point x = (t,y) E U the set F(G) be bounded. Then for every neighborhood V of the set cont,(D(F), x) and for every c > 0 the set G-(V, c) = {x} U {(t*,y*): t* E (t - e,t), y* E Uto, ~::::r E V} is a left D(F)-neighborhood of the point x. • Analogous remarks and theorems are true for right D(f)-neighborhoods, right contingents and the set G+(V, c) = {x} U {(t*, y*): t* E (t, t + c), y* E Uto, ~::::r E V} and for D(F)-neighborhoods, contingents and the set
G(V, c) = G-(V, c) U G+(V, c). Example 5.1. Consider the scalar differential equation I
Y
=
y2 t
2
+y 2
+ t.
Peano, Caratheodory and Davy conditions.
269
Its right hand side is defined on the entire plane except the point (0,0). The solution space Z of the equation does not depend on the definition of the right hand side at the point (0,0). Define the right hand side at the point by putting 1(0,0) = O. The function 1 is continuous on the entire plane except the point (0,0). In order to use Theorem 5.1 in investigation of our equation we will establish the continuity of the function 1 with respect to the topology rZ. A crude estimate of the right hand side gives the inequality ly'(t)l:::;; 1 + Itl. Therefore cont(Z,(O,O)) ~ [-l,lJ. Let a = inf{a: a;?: 0, cont(Z,(O,O)) ~ [-a, a]}. Evidently cont(Z,(O,O)) ~ [-a,aJ. Show that a = 0. Assume the opposite. Take an arbitrary 6 > 0 and apply Theorem 5.2 to the set V = (-a - 6, a + 6). For every point (t, y) (f (0,0)) of the set G(V, c) we have
Ity I :::;;a+6,
2
y2
t ;?: (a+6)2' ly'(t)l:::;;x(6,c),
1)) (a+<5)2 h were x ( u, c = 1+(a+6)2 + c. The last estimate implies that cont(Z, (0, 0)) this is true for every 6 > 0 and c > 0,
~
a2
a2
[-x(6, c), x(6, c)J. Since
cont(Z,(O,O))~ [ -1+a 2 '1+a 2
] •
Since 0 < a :::;; 1, then 1~:2 < a2 :::;; a. The obtained inclusion contradicts the definition of the number a. Thus our assumption is false and hence a = O. By Theorem 5.2 for 6 > 0 and c > the set G(( -6,6), c), is a Z-neighborhood of the point (0,0). For every point (t, y) of this set we have II(t,y)1 :::;; 1!~2 + c. This implies the continuity of the function 1 with respect to the topology rZ at the point (0,0). By remarks of the first paragraph of our reasoning we have the continuity on the entire plane. By Theorem 5.1 all solutions of our equation are continuously differentiable.
°
6. General approach to existence theorems
~
This section does not contain new essential information in comparison with the accounted material and its contents has a character of a comment. But this comment cost to give it. In textbooks on the theory of Ordinary Differential Equations authors pay more attention to theorems on the existence of solutions of the Cauchy problem than to theorems on the continuity of the dependence of solutions on parameters. Psychologically this is easy to understand: when an existence theorem is not proved then what is the continuity we must discuss?
CHAPTER 8
270
Continuity theorems play the role of a comment to existence theorems. The existence theorems take a central place. However, contrary to a continuing belief in an absolutely greater importance of the existence in comparison with the continuity, in reality the existence is a secondary property. We establish it easily each time when a corresponding result on the continuity in the form of the convergence of corresponding sequences of solution spaces is obtained. When the convergence result is proved it remains to use Theorem 7.2.1. We had such situations above. As a simple exercise the reader can deduce the Peano theorem from the Picard and Weierstrass theorems. Let U be an open subset of the product ~ x ~n. For an equation with the continuous right hand side the Peano theorem and remarks of §2 and of Example 6.4.1 give the following result. Theorem 6.1. Let a mapping f : U --+ ~n be continuous. Then D(J) E Rcek(U), Following completely just mentioned scheme remarks of §2 expand the class of equations and inclusion considered in Theorem 6.1 to the class of differential inclusions y' E F(t, y) with right hand sides F E Qd(
Pk(X)=
o k(P(x,E)-~) 1
for p(x, E) < for
1
k~
1
k'
p(x,E) ~
for p(x, E)
2
k'
2
> k'
The function gk(t, y) = Pk(t, y)f(t, y) is continuous. By Theorem 6.1 D(gk) E Rcek(V), The convergence D(gk, U) --+ D(J, U) is obvious: for ev~ ery point x E U there exist its neighborhood Ox and an indeX ko = 1,2, ... , such that gk lox = fl ox for k = ko, ko + 1, ko + 2, .... NoW
Peano, Caratheodory and Davy conditions. Theorem 6.6.2 implies the convergence D(9k) D(f) E Rcek(V).
---+
271
D(f). By Theorem 2.1
7. Majorants of right hand sides The proof of Theorem 2.1 used properties of the solution space of the inequality Ily'(t)11 ~ cp(t) investigated in §6.3. In this section we will try to expand the list of differential inclusions, which may be used as majorants in results as Theorem 2.1 (now this list consists only of the mentioned inequality). We will try to understand how we may prove that the solution space of a differential inclusion possesses the corresponding properties. As before U denotes an open subset of the product ]R x ]Rn. Denote by Q**(U) the set of all multi-valued mappings F E Q*(U) (see §6.4) satisfying the condition: for every y E ]Rn the mapping FY is measurable. The last condition is near to condition a) of §2 (see Theorem 4.13.3). The condition FE Q*(U) is near to condition b) of §2 (see remarks of §5.4). Denote by Q;*(U) the set of all mappings F E Q**(U) with compact values. For (arbitrary) multi-valued mapping G : U ---+ ]Rn denote by Q*c (U) the set of all multi-valued mappings F E Q*(U) satisfying the conditions:
(7.1) F(x)
~
G(x) for every x E Uj
(7.2) if x E U and G(x)
1= 0,
then F(x)
1= 0.
Put Q**C(U) = Q*C(U) n Q**(U), Q;*C(U) = Q;*(U) n Q**C(U). When we estimate the relation between F and G in the last definitions we can word condition (7.1) in the following way: the mapping G majorizes (is a majorant of) F. Following the way to our aim, select a set of all suitable majorants and introduce the set Qme(U) of all multi-valued mappings G : U ---+ ]Rn satisfying the conditions
(7.4) D(F) E Re(U) for every mapping FE Q;*C(U). With this notations the problem consists in the proof of the membership G E Qme(U). We will need the following auxiliary assertion. Lemma 7.1. Let -00 ~ a < b ~ 00, U ~ (a,b) x ]Rn, G E Q*(U). Let Ip : (a,b) ---+ (0,00). Let H(t,y) = O,(G(t,y),cp(t» for (t,y) E U. Then II E Q*(U).
CHAPTER 8
272
Proof. Let (t, y) E U. For every neighborhood V of the point y in Ut and for every c > 0 we have
Ht(V) = U{O,(G(t,y*),cp(t)): y* E V} ~ U{O(G(t,y*),cp(t) +c): y* E V} = O(Gt(V), cp(t) + c), c(Ht(V)) ~ c(O(Gt(V), cp(t) + c)) ~ O(c(Gt(V)), cp(t) + c) ~ O(cc(Gt(V)), cp(t) + c) (see Lemma 2.4.3),
If a point p belongs to the set n{ cc(Ht(V)): V is a neighborhood of the point y} then for every neighborhood Oy of the point y the set
O(p, cp(t)
+ 2c) n cc(Gt(Oy))
~
O,(p, cp(t)
+ 2c) n cc(Gt(Oy))
is nonempty. Therefore, the family {O,(p, cp(t) + 2c) n cc(Gt(V)): V is a neighborhood of the point y} of compact subsets of the set 0, (p, cp( t) + 2E) has finite intersection property. By Theorem 1.6.3 its intersection Me is nonempty. By the same arguments the set
M
= n{Me:
c> O} ~ n{O,(p, cp(t)
+ 2c) : c >
O}
= O,(p, cp(t))
is nonempty too (it lies in the compactum n{cc(Gt(V)): V is a neighborhood of the point y} ~ Gt(y) ). Hence p E Ht(y). In view of the arbitrariness in the choice of the point p this means, that n{cc(Ht(V)): V is a neighborhood of the point y} ~ Ht(y). In view of the arbitrariness in the choice • of the point (t, y) E U this implies our assertion. Lemma 7.2. Let under hypotheses of Lemma 7.1 G E Q**(U) and the function cp be measurable. Then H E Q**(U). Proof. The proof reduces to referring to Lemmas 7.1 and 4.12.3. • Lemma 7.3. Let F1,F2 E Q*(U). Let H(t,y) = F1(t,y) n F2 (t,y) for (t, y) E U. Then H E Q*(U). Proof. Let (t, y) E U. For i = 1,2 we have n{ cc{Ht(V)): V is a neighborhood of the point y} ~ n{CC{{Fi)t{V)): V is a neighborhood of the point y} ~ Fi(t, y). Therefore n{ cc({H)t{V)): V is a neighborhood of the point y} ~ Fl (t, y) n F2 {t, y) = H{t, y). In view of the arbitrariness in the choice of the point (t, y) E U this • implies our assertion.
273
Peano, Caratheodory and Davy conditions.
Lemma 7.4. Let under hypotheses of Lemma 7.3 F I , F2 Then H E Q**(U).
E
Q**(U).
Proof. The proof reduces to referring to Lemmas 7.3 and 4.12.4. • Lemma 7.S. Let G E Q**(U), D(G) E Rcn(U), Go,F E Q**G(U), Q;*Go(U) i= 0. Let for every point x E U either F(x) = G(x), or the set F(x) be nonempty and compact. For r > 0 define the mapping G r : U ---7 ~n, (7.5)
Then D(F) E [U{D(A): A E Q;*Gr(U)}: r ~ O}]Rc(U) n Rcn(U). Proof. I. Since D(F,K) = D(F) nD(G,K) for every compactum K ~ U, by Theorem 6.4.2 and Lemma 6.4.4 D(F) E Rc(U). The membership D(F) E Rn(U) is obvious. Thus the problem consists in the proof of the membership D(F) E [U{D(A): A E Q;*Gr(U)} : r ~ O}]Rc(U). Do it.
II. Fix a function A E Q;*Go(U). For x E U and k = 1,2, ... put n Of(A(x), k). Remark 2.3.3 and Theorem 2.4.1 imply the compactness of values of the mapping Ak : U ---7 ~n. By Lemmas 7.2 and 7.4 Ak E Q;*Gk(U). III. Put Mk = {x: x E U, F(x) U A(x) ~ 0(0, k), G(x) i= 0} for k = 1,2, .... By Theorem 4.12.1 for every y E ~n the set (Mk)Y is measurable. Theorems 2.5.2 and 1.7.5 imply that for every t E ~ the set (Mk)t is open (in the space ~n). For x E Mk the triangle inequality implies that F(x) ~ A2k(X). Therefore the mapping Bk : U ---7 ~n,
Ak(X) = G(x)
for x E M k , for x E U\ M k , belongs to Q~*G2k (U). IV. Show that the sequence a = {D(Bk): k = 1,2, ... } converges in U to the space D(F). Since all elements of the sequence a lie in D(G), the condition D(G) E Rc(U) implies the membership a E s(U). By Lemma 7.1.4 our aim will be achieved when we prove the inclusion llin top SUPk-+oo D(Bk ) ~ D(F). Let Z E lim topsUPk-+oo D(Bk). If 7r(z) consists of one point only, then Z E D(F) by definitions. Let the domain [a, b] of Z contain more than one point (i.e., a < b). By Lemma 1.5.1 we can choose functions Zj E D(BkJ for j = 1,2 ... (where kl < k2 < k3 < ... ) such that the sequence {Zj : i = 1,2, ... } converges to z. Let I be an arbitrary segment lying in (a, b). By virtue of the continuity of the mapping 7r we have I ~ 7r(Zj), beginning with some j = jo. By the Condition D(G) E Rn(U) for every point t of a subset P of full measure of
274
CHAPTER 8
the I the derivatives z'(t) and zj(t), j = jo,jo + l,jo + 2, ... , exist, belong to the sets G(t, z(t)) and G(t, Zj(t)), respectively, and z'(t) E n{ cc{ zj(t): j = j* ,j* + 1, ... } : j* = jo,jo + 1, ... }. For t E P we have two possibilities. 1. F(t, z(t)) = G(t, z(t)). Then z'(t) E F(t, z(t)) by the choice of the set P. 2. The set F(t, z(t)) is nonempty and compact. In this case (see Lemma 2.3.6) for some ko = 1,2, ... the point z(t) belongs to the open subset V = (Mko)t of the space ~n and V ~ (Mdt for k = ko + 1, ko + 2, .... So Bk; (t, y) = F(t, y) for y E V and kj ~ ko. The convergence Zj(t) --t z(t) and the condition F E Q*(U) imply that
z'{t) En {cC{{U{Bk;(t,Zj(t)) : j =j*,j* + 1,j* + 2, ... , kj ~ F{t, z(t)). We have proved that Z
ZII
~
ko} : j* = 1,2, ... }
E D(F). By Corollary of Lemma 6.2.1
E D(F).
V. Referring to Theorem 7.2.1 completes the proof. • A mapping Go E Q**G(U) is called a sufficient me-approximation of a mapping G : U --t ~n, if for every r > 0 the mapping G T : U --t ~n defined by (7.5) belongs to Qme(U). Theorem 7.1. Let a mapping G E Q**{U) possess a sufficient me-approximation Go and D(G) E Rcn(U). Let F E Q**G(U). Let for x E U either F(x) = G(x), or the set F(x) be nonempty and compact. Then D{F) E Rcen(U). Proof. The proof reduces to referring to Lemmas 7.5 and 7.7.2 and Theorem 7.7.1 and 7.2.1. • Corollary. Under hypotheses of Theorem 7.1 G E Qme(U) and D(G) E Rcen(U). • Denote by Qmek(U) the set of all multi-valued mappings G E Qme(U) satisfying the condition
(7.6) D(F) E Rk(U) for every mapping F E Q;*G(U). A mapping Go E Q**G(U) is called a sufficient mek-approximation of a mapping G : U --t ~n, if for every r > 0 the mapping G T : U --t ~n defined by (7.5) belongs to Qmek(U), Theorem 7.2. Let under hypotheses of Theorem 7.1 the mapping Go be a sufficient mek-approximation of the mapping G. Then D(F) E Rcekn(U). Proof. The proof is analogous to the proof of Theorem 7.1. • Corollary. Under hypotheses of Theorem 7.2 G E Qmek(U) and D(G) E Rcenk(U), •
Peano, Caratheodory and Davy conditions.
275
Lemma 7.6. Let -00 :::;; a < b :::;; 00, U ~ (a, b) x IRn, G E Q**(U). Let a function cp : (a, b) --+ (0, (0) be locally Lebesgue integrable. Let the set Go(t, y) = G(t, y) n 0,((5, cp(t)) be nonempty for every point (t, y) E U. Then the mapping Go : U --+ IRn is a sufficient mek-approximation of the mapping G. Proof. By Lemma 7.4 Go E Q**(U). By Theorem 4.12.3 the mapping Go satisfies the Davy conditions (with cp as majorant). Lemmas 7.2 and 7.4 imply that for every r > 0 the mapping G r : U --+ IRn defined by (7.5) satisfies the Davy conditions. This gives what was required. The lemma is proved. • Theorem 7.3. Let under the hypotheses of Lemma 7.6 D(G) E Rcn(U). Then G E Qmek(U). Proof. The proof reduces to referring to Lemma 7.6 and Corollary of • Theorem 7.2. Example 7.1. Let U = IR x R For (t, y) E U denote by G(t, y) the segment [1,1 + Iyl-l], if y =F 0, and the half interval [1,(0), if y = 0. For (t, y) E U put h( t, y) = y. The line y = 0 is finite with respect to the space M1(h, 1) (see §7.4), Ml(h, l) ;2 D(G). By Theorem 2.1 outside this line the space D(G) satisfies conditions (c) and (n). Theorem 7.6.6 and Lemma 7.4.2 implies that D(F) E Rc(U). By Theorem 7.4.3 and Lemma 3.1 D(F) E Rn(U). By Theorem 7.3 G E Qmek(U). Theorem 7.4. Let {G k : k = 1,2, ... } ~ Qme(U) (respectively, ~ Qmek(U)), D(Gd E Rcn(U). Let G1(x) ;2 G 2 (x) ;2 G 3 (x) ;2 ... and k = 1,2, ... } for every point x E U. Then G(x) = n{Gk(x): D(F) E Rcen(U) (respectively, D(F) E Rcekn(U)) for every mapping FE Q;*G(U). Proof. For k = 1,2, . .. and x E U put if F(x) = 0 if F(x) =F 0. Evidently Fk E Q**(U). By Theorem 7.1 D(Fk ) E Rce(U) (respectively, by Theorem 7.2 D(Fk ) E Rcek(U)). Since for every point x E U we have F(x) = n{Fk(X): k = 1,2, ... }, by Lemmas 2.2 and 2.3 the sequence {D(Fk): k = 1,2, ... } converges in U to D(F). By Theorem 7.2.1 (respectively, by Theorem 2.1) D(F) E Rce(U) (respectively, D(F) E Rcek(U)). The fulfilment of condition (n) is obvious. The theorem is proved. • Corollary. Let under hypotheses of Theorem 7.4 Q;*G(U) =F 0. Then G E Qme(U) (respectively, G E Qmek(U)). • Example 7.2. Let U = IR x JR. For (t, y) E U denote by G(t, y) the segment [1 + Iyl-l, 1 + 2Iyl-l] if y =F 0, and the empty set if y = o.
276
CHAPTER 8
For k = 1,2, ... and (t, y) E U denote by Gk(t, y) the segment G(t, y), if Iyl ~ k- 1 , the segment [1 + k, 1 + 2Iyl-l], if 0 < Iyl ~ k- 1 and the half interval [1 + k, 00), if y = O. When we repeat reasonings of Example 7.1 we show that G k E Qmek(U) for every k = 1,2, .... We have G1(x) 2 G 2 (x) 2 G 3 (x) 2 ... and G(x) = n{ Gk(x): k = 1,2, ... } for every point x E U. By Corollary of Theorem 7.4 G E Qmek(U),
CHAPTER 9
COMPARISON THEOREM
Explicit aim of this chapter is to expand results related to differential inequalities. They allow us to compare the behavior of solutions of two ordinary differential equations. The position of such assertions is well known. In comparison with the existing results on this subject we weaken essentially restrictions on the right hand side of the majorizing equation. In particular, we will consider equations with the multi-valued right hand sides (differential inclusions). Here in restrictions on the right hand side we move a little outside the framework of Davy (Caratheodory) conditions. The implicit, but main, aim of this chapter is to show how we can investigate discontinuous right hand sides of equations and inclusions in complicated situations (related here to differential inequalities). The author does not insist that the multidimensional version of the comparison theorem proposed here is the best of possible ones. On the contrary, in comparison with the corresponding theorem for equations with continuous right hand sides the proposed assertion looks cumbersome. It may be that this will stimulate the appearance of a new more 'understandable' assertion on this subject. On the other hand, in the statement of this result we do not go outside the framework of the main general ideas of our approach, and when the reader has any initial experience of insight within this circle of notions the mentioned assertion does not look so terrible. The scheme for the passage from the scalar case to the multidimensional one in the proof of the corresponding classical result (see, for instance, [RHL]) keeps its value for equation with a discontinuous right hand side too. We shall not discuss this question here. On the other hand, this simple method does not work in some situations, because even the addition of a constant term in the right hand side may lead to the necessity of proving anew the presence of basic properties of solution spaces (as (c), (e), etc.). Our method (more cumbersome in the statement and proof of the main result) may turn out to be simpler in the investigation of particular equations. Our case of the comparison theorem for the scalar case also does not represent its final version. Although here we make an essential step in weakening restrictions imposed on the right hand side, we have possibilities of
278
CHAPTER 9
moving further in this direction. But the employment of these possibilities will need further effort (and complications of statement and proofs). 1. Existence of upper solutions in the scalar case
Let U be an open subset of the plane ~ x R Theorem 1.1. Let Z E Rce(U) and (to, Yo) E U. Then there exists a function Zo E Z+ such that to = inf7r(zo), zo(to) = Yo and:
(1.1) if Z E Z, tl = inf 7r(z) E 7r(zo) and z(t 1 ) for every t E 7r(z) n 7r(zo) (Figure 9.1).
~
zo(td, then z(t)
~
zo(t)
Proof. I. There exists a number v > 0 such that
[to - v, to
+ v]
~
7r(z)
for every solution of the Cauchy problem y E Z-+, y(t o) = Yo, see Lemma 6.6.3. II. For t E [to, to + v] put a(t) = sup{z(t): Z E Z,
u i
+-----ro---
+ v], z(t o) = Yo}. tI Show that a E Z. Let {t k : k = 1,2, ... } be a Figure 9.1 dense subset of the segment [to, to + v]. For every k = 1,2,... fix a function Zk E Z such that 7r(Zk) = [to, to + v], Zk(t O) = Yo and zk(td = a(tk). Such a function exists because condition (c) implies that:
7r(z) = [to, to
a(tk) E {z(t k ): Z E Z, 7r(z) = [to, to Construct a sequence of functions {ak: k = 1,2, ... } ~ Z. Put al = Zl· Let functions al, ... , ak-l be pointed. Fix a function ak. By the definition of a( td we have: ak-l(t k ) ~ a(t k). If ak-l (t k) a(t k ), we put ak = ak-l· If ak-l(tk) < a(tk), then the open set M = {t :
t
E
[to, to
+ v],
z(t o) = Yo}.
tk "" -- .... _-_._-+.--=----..
J
+ v], ak-l (t) < Zk(t)}
is a neighborhood of the point t k , see Figure 9.2.
Figure 9.2
279
Comparison theorem.
Let J :3 tk be a connected component of the set M. The function: for for is defined on the segment [to, to
+ v).
t E J,
t E [to, to
+ v) \
J
It belongs to Z and
(1.2) Because of the membership Z E Rc (U) the sequence {ak : k = 1, 2, ... } contains a subsequence converging to a function (3 E Z, 7r({3) = [to, to + v). By (1.2) (3(t k ) = a(tk) for every k = 1,2, .... Show that (3(t) = a(t) for every t E [to, to + v). The definition of a and the membership (3 E Z imply the estimate a(t) ~ (3(t). Thus it remains to obtain the estimate a{t) ~ (3(t). If this is false then there exists a function Z E Z such that z{t) > (3(t). By virtue of the density of the set {tk: k = 1,2, ... } in the segment [to, to + v) we have z(t k ) > (3(t k) for some k = 1,2, .... As we have established above, this is false. Thus (3 = a, which gives what was required. III. The facts proved in I-II may be summed up in the following way: there exist a number v > 0 and a function a E Z such that: (1.3) (1.4)
7r(a)
= [to, to + v) and a(t o) = Yo;
if Z E Z, 7r(z) ~ [to, to + v) and z(inf 7r(z)) ~ a(inf 7r(z)), then z(t) ~ a(t) for every t E 7r(z).
In fact, if the function a E Z from II does not satisfy the condition (1.4) then there is a function Zl E Z such that 7r(zd ~ [to, to some tl E 7r(zd.
+ v),
zl(inf7r(zd) ~ a(inf7r(zd), and zl(td > a(td for
The continuous realfunction 6(t) = a(t)-zl (t) is defined on the segment 7r(zd· It must vanish at a point s on the segment [inf 7r(zd, td. The function
Z2 (t ) -_ {a(t) zdt)
for for
E
[to, s)
t E
t
Is, t l )
belongs to Z, to E 7r(Z2) and Z2(t O ) = a(t o) = Yo. By the definition of v in I the function Z has an extension Z3 E Z defined on the segment [to, to + v]. We have z3(td > a(td, which contradicts the definition of a(td in II and gives what was required.
280
CHAPTER 9
IV. If for i = 1, 2 the pair consisting of the function ai (as a) and of [t ]. the number 1/i (as 1/) satisfies (1.3 - 4) and 1/1 ~ 1/2, then al = 0,111 Therefore the condition:
a21
to = inf1r'(z) and for every
1/
< SUP7r(z) the function
zl[to,v]
satisfies
(1.3-4) defines an unique function z E Z+. It satisfies all imposed conditions. The • theorem is proved.
2. Scorza-Dragoni property We will consider a differential inclusion, the right hand side F of which is defined on an open set U ~ ~ X ~n, moreover: (2.1) values of the mapping Fare nonempty compact and convex; (2.2) for every fixed t E
~
the mapping F(t, y) is upper semicontinuous in
y;
(2.3) if a compactum K ~ ~n and a segment I ~ ~ are such that IxK then for every open subset V of the space ~n the set
~
U,
M*(K, V) = {t: tEl, F({t} x K) ~ V}
is measurable. Compare conditions (2.1-3) with remarks of §§4.12-13 and 8.2. We say that a differential inclusion y' E F( t, y) satisfies the ScorzaDragoni condition if for every point x E U it has a neighborhood Ox ~ U and a subset M of the measure zero of the real line such that: if z E D(F, Ox) and t E (7r(z») \ M, then all limit points of the quotient z(t
(2.4) as h
+ h) -
z(t)
h --t
0 belong to F(t, z(t»).
(It seems that the first person who paid attention to properties of this kind was Scorza-Dragoni, [Sc2].) Lemma 2.1. Let conditions (2.1 - 3) hold. Assume that (2.5) there exists an integrable function cp : lR for (t, y) E U and u E F(t, y).
--t
[0,00) such that
lIull
~ cp(t)
281
Comparison theorem.
Then the differential inclusion y' E F(t, y) satisfies the Scorza-Dragoni condition. Proof. 1. Take an arbitrary x E U. Find a segment / ~ R and a bounded open set G ~ Rn such that
x E (/) x G
~
/ x [G] ~ U.
II. Let H be a closed subset of the segment / and X be the characteristic function of the set / \ H. For almost all t E H t+h
J
lim -hI
(2.6)
h--.O
X(s)
t
Let Q be the set of all density points t of the set H satisfying (2.6). Let the mapping FIHX[G] be upper semicontinuous. III. Let with the notation of I, II z E D(F), t E (7r(z») n Q and (t, z{t)) E (/) x G. Take arbitrary 8 > O. By virtue of the upper semicontinuity of the mapping F on the set H x [G] for some neighborhood W ~ (/) x G of the point (t, z(t») we have
F(W S
n (H
x [G])) ~ O/jF(t, z(t».
Let a > 0 be such that (t - a, t E (t - a, t + a). IV. Let hE (0, a). Then
J
=
h1
~
7r(z) and (s, z(s» E W for all
J t+h
Hh
z(t+h)-z(t) h
+ a)
z'(s)ds
h1
=
t
J
'
Hh
1 x(s)z'(s)ds+ h
(I-X(s»z (s)ds.
t
t
By (2.6) t+h
Hh
lJ
X(s)z'(s)ds ::;;
lJ
t
X(s)
-t
0
as
h
-t
O.
t
Take an arbitrary
U
E
f(s) =
F(t, z(t» and consider the function
{~(S)
for for
7r(z) n H, s E 7r(z) \ H. S
E
All its values belong to O/jF(t, z(t». It is Lebesgue integrable. Therefore its mean value Hh
lJ t
f(s)ds
CHAPTER 9
282
belongs to [06F(t, z(t))], see Theorem 4.10.1. In addition,
~
t+h
J t
f(s)ds -
~
Hh
*J
Hh
J
(1 - X(s))z'(s)ds =
t
X(s)ds
-7
0
as
h
-7
0,
t
because t is a density point of the set H. Thus limit values of the expression
~
Hh
J
(1 - X(s))z'(s)ds,
t
and limit values of the quotient (2.4) as h -7 0 and h > 0 belong to [06F(t, z(t))]. Since this is true for every 8 > 0, limit values of the quotient (2.4) as h -7 0 and h> 0 belong to F(t, z(t)). V. In complete analogy with IV we establish that limit values of the quotient (2.4) as h -7 0 and h < 0 belong to F(t, z(t)). VI. Theorem 4.13.2, VI and V imply what was required. The lemma is proved. • Lemma 2.2. Let conditions (2.1- 3) hold. Let H ~ lR be a closed set of measure zero. Let
Let U be an open subset of the plane lR x R Theorem 3.1. Let F : U -7 lR be multi-valued mapping with connected values. Let the differential inclusion y' E F(t, y) satisfy the Scorza-Dragoni condition, Z = D(F) E Rce(U). Let a function Zo E Cs(U) l'e generalized absolutely continuous and z~(t) ~ supF(t, zo(t)) for almost all t E 7r(zo)· Let Zl E Z+ be a solution of the differential inclusion y' E F( t, y), the existence of which is established in Theorem 1.1 with to = inf 7r(Zl) and 7r(zo) ~ 7r(zd, Yo = Zl(tO) ~ zo(to)· Then zo(t) ~ Zl(t) for all t E 7r(zo). Proof. The inequality
(3.1) is true at least for t = to. Evidently by virtue of the connectedness of the segment 7r(zo) in order to prove the theorem it is sufficient to check that the set M of points t E 7r(zo), for which the inequality (3.1) holds, is open
Comparison theorem.
283
in 7f(zo) (its closedness is obvious: M = 'ljJ-l([O, 00)), where 'ljJ = Zl - Zo, see Theorem 1.7.5). Because of the possibility of changing the value of to it is sufficient to prove the existence of E > 0 such that the inequality (3.1) holds for all points t E 7f(zo) n [to, to + E). Fix E > 0 such that every solution, the graph of which meets the graph of the function zl[to,to+el' is defined at least on the segment [to, to + Ej. A possibility of such a choice of E > 0 follows easily from Lemma 6.6.3. The set of points t E [to, to + E] satisfying the condition: (3.2) there exists a function z(inf7f(z)) = zo(inf7f(z)), z(s)
Z
>
E Z such that to ~ inf7f(z) ~ zo(s) for all s E (inf7f(z),t]
t E 7f(z),
is open in the segment [to, to + E] and falls into the union of the family, of its connected components. Let J E ,. Let ZJ E (D(F))+ be an upper solution of the inclusion y' E F(t,y) (in the sense of the fulfilment of (1.1)) with zJ(inf J) = zo(inf J). Because of the definition of J and (3.2) zJ(t) > zo(t) for every t E J. If in addition sup J f. to + E, then zJ(sup J) = zo(sup J). Define a sequence of function
t E [to, to t E Jk •
+ E] \ J k,
If the family , is finite our construction will end on a finite step k by the pointing of a function
supF(t,
and
(3.4) t E J E ,.
CHAPTER 9
284
Thus the function cp satisfies all conditions imposed on the function Zo. The definition of the function cp implies also that
(3.5)
cp(t)
~
zo(t)
for all
t
E
[to, to
+ e]
and (3.6) if z E Z, inf7r(z) E [to, to + e] and z(inf7r(z)) = cp(inf7r(z)), then z(t) ::;; cp(t) for all t E 7r(z) n [to, to + e]. From the Scorza-Dragoni condition and the generalized absolute continuity of the function cp there exists a subset Q of full measure of the segment [to, to + e], for every point t of which the approximate derivative cp'(t) exists, limit values of the quotient Z(Hhl-z(t) as h - t 0 lie in F(t, cp(t)) = F(t, z(t)) for every solution Z of the Cauchy problem Z E Z, t = inf(7r(z)) and z(t) = cp(t). By (3.6)
z(t + h~ - z(t) ::;; cp(t + h~ - cp(t)
(for
h
> 0).
Therefore cp'(t) ~ inf F(t, cp(t)). When we compare this fact with (3.3-4), we obtain: cp'(t) E F(t, cp(t)) for all t E Q. SO cp E D(F). By the definition of Zl we have cp(t) ::;; Zl(t) for t E [to, to +r:]. By (3.5) zo(t) ::;; Zl(t) for t E [to, to + e] and [to, to + e] ~ M. The theorem is proved. •
4. Comparison theorem: n-dimensional case Assume now that U = T X U1 X ..• X Un is an open subset of the product JR. x JR.n and T, U1 , ••• , Un are intervals of the real line. Following the way outlined in the previous section we pass here from general hypotheses of comparison theorems to the consideration of multivalued mappings F : U - t JR.n, such that
(4.1) F(t, Yl,' .. , Yn) mappings Fi : U - t JR.,
= Fl (t, Yl,' .. , Yn) x· .. x Fn(t, Yl, ... , Yn),
where the i = 1, ... , n, satisfy (2.1-3) (with the change n to
1), (4.2) supFi(t,XI,""X n)::;; supFi(t,YI, ... ,Yn) for all values of arguments from U satisfying the condition Xl
~ YI,""
Xi-l
~
Yi-l, Xi = Yi, Xi+! i
~
Yi+!,···, Xn
~
Yn'
285
Comparison theorem.
In the case where each mapping F i , i = 1, ...
,n,
does not depend on
Yl,'·" Yi-ll Yi+l," . ,Yn condition (4.2) may be omited: here the inclusion y' E F(t, y) falls into the system of independent scalar inclusions:
(4.3)
Results of the previous section are applicable to it (in every coordinate apart). This implies the existence of an upper (in the sense of (1.1)) solution of (4.3) with respect to the partial order on IRn , which is usualy considered in connection with the questions under concideration (see [RHL]):
Besides the mapping F consider the mapping
where each mapping G i : T X Ui -+ IR n , i = 1, ... ,n, satisfies the conditions imposed in Theorem 3.1 on the mapping F. For a function Z = (Zl' ... ,zn) E (D( G) )-+ and for i = 1,2, ... ,n put
and
wt(t, y) = Fi(t, Zl (t), ... ,Zi-l (t), y, Zi+l (t), ... ,zn(t)). Assume that:
(4.6) for every function Z E (D(G))-+ the right hand sides of the differential inclusions y' E
286
CHAPTER 9
(4.7) (to,yo) E U, Yo = (Yl, .. ·,Yn) and c > 0 be such that for evand ery solution of the Cauchy problem Z E (D(C))+, inf7r(z) = to z(t o) = Yo we have to + c < SUP7r(z) (see Lemma 7.6.3 implying the existence of such c for an arbitrary point (to, Yo) E U). Let Z E D(C), to E 7r(z) and z(t o) = Yo. By Theorem 1.1 for every i = 1, ... , n there exists an upper solution Zi of the Cauchy problem y' E cpt(t,y), y(to) = Yi (see the remark about the system (4.3) ). Denote O(z) = (Zl,'" ,zn)I[to,to+c] E D(cpZ). A function z E (D(C))-+, 7r(z) = [to, to + c], will be called correct, if for every t E [to, to + c] the value of the function O(z) at t is not greater than the corresponding value of the function Z (with respect to the above partial order, we will write: O(z) -< z). Remark 4.1. If Zl -< Z2, then O(zd -< 0(Z2)' This is obvious. This implies that if a function Z is correct then the function O(z) is correct too (that is, O(O(z)) -< O(z)). Remark 4.2. Correct functions exist. An upper (for t ~ to) solution of the Cauchy problem y' E D(C), y(t o) = Yo is such a function. Remark 4.3. If a sequence Zl >- Z2 >- Z3 >- ... of correct functions converges uniformly to a function Z then by Remark 4.1 O(z) -< O(Zi) -< Zi for every i = 1,2, .... Therefore O(z) -< z. That is, the function Z is correct. Now we can define a transfinite sequence of functions {z,,: a < wd, (4.8)
in the following way. Take as Zl the correct function from Remark 4.2. Define the function Z"+1 by the formula Za+1 = O(za)' For a limit transfinite a = .lim ai, al < a2 < a3 < ... , we do as follows. By virtue ofthe condition '-+00 D(C) E Rce(U) the sequence {za o : i = 1,2, ... } contains a subsequence converging uniformly on the segment [to, to + c] to a function Z E D(C). Now (4.8) implies the uniform convergence z{3 --t Z as f3 --t a. By Remark 4.3 the function Z is correct. Put Za = z. Functions z" are ordered in each coordinate. This implies easily, that Za = Za+1 = Za+2 = ... , beginning with some a. So there exists a function ~ = (~l"" '~n) E D(C), such that 0(0 =~, (7r(O = [to, to + cD. By the definition of 0 and cpz
(4.9)
~Ht) ~
sup Fi(t, ~l (t), ... '~n(t)) for almost all t E [to, to
Remark 4.4. Let (4.10)
+ c].
Comparison theorem.
287
By (4.2) and by the definition of <1.>z for every correct function Z >- Z· the function z· solves the inclusion Y' E <1.>Z(t,y). As O(z) is an upper solution of the last differential inclusion, then O(z) >- z·. Remark 4.4 implies easily that the above function ~ majorizes all solutions of the Cauchy problem (4.10). Show that, in fact, (4.11) Consider the inclusion Y' E
(4.12)
wf(t, y).
By Theorem 1.1 there exists an upper solution "1 of the Cauchy problem for (4.12) with the initial value y(t o) = Yi. By virtue of the equality
and of the construction of the upper solutions ~i' "1, and by Theorem 1.1, "1. This implies the fulfilment of (4.11). Remark 4.5. Let now [t I , t2J ~ [to, to + cJ. Let a function
~i =
satisfy the conditions ((t I ) -<
~(td
and
(4.13) (Ht) ~ SUpFi(t'(I(t), ... ,(n(t)) for i = 1, ... ,n and for almost all
t E [t I , t2J. By (4.13) the function ( satisfies the condition:
By Theorem 3.1 Show that: (4.14)
( -<
ZI.
( -<
Za
for every
a <
WI.
Let the assertion hold for some a. Show that it holds for a + 1 too. By (4.2) (:(t) ~ sup Fi(t, YI (t), ... , Yi-I (t), (i(t), YHI (t), ... , Yn(t)), where (YI, ... , Yn) = Za· Therefore (:(t) ~ sup <1.>:'" (t, (i(t)) for almost all
t E [t l , t2J. By Theorem 3.1 (-< O(zaJ = Za+!. The passage over a limit transfinite a in (4.14) is obvious.
288
CHAPTER 9
Thus (4.14) is proved completely. For some a the function z'" coinsides with the upper solution ( Thus we have proved: Theorem 4.1. Let conditions (4.1 - 2) hold. Let condition (4.5) hold for a mapping G : U ----> IRn,
with the mappings G i : Ui ----> IR satisfying (2.1 - 3). Let the differential inclusions y' E G(t, y), y' E Z(t, y) and y' E \lIf(t, y) satisfy the ScorzaDragoni condition and D(G) E Rce(U), D(Z) E Rce(U), D(\lIf) E Rce(U) for all z E D(G) and i = 1, ... , n. Let condition (4.7) hold. Then there exists a solution Zo of the Cauchy problem y' E F(t, y), y(t o) = Yo on the segment [to, to + c] such that: if a generalized absolutely continuous function z E Cs(U) is defined on a segment [t l , t 2 ] ~ [to, to + c] and satisfies the condition z(td -< zo(t l )
and
z'(t) -< supF(t, z(t))
for almost all t E [tl' t 2 ],
then z -< Zo (on [tl, t 2 ]J.
•
(Here and below in the next {sup FI (t, z(t)), . .. , sup Fn(t, z(t))}.)
section
supF(t,z(t))
denotes
5. Localness of the property in question When we passed from the scalar case to the multidimensional case we introduced an auxiliary differential inclusion (5.1 )
y' E G(t, y)
majorizing the inclusion (5.2)
y'
E
F(t, y)
under consideration. The pointing of an auxiliary inclusion (5.1) on the entire domain of the right hand side of (5.2) may have obstacles of a purely technical character which may be easily passed over. This is analogous to the case of a linear differential equation with nonzero coefficients. It has no Caratheodory majorant (in the entire domain of the right hand side), although locally this majorant may be pointed in a completely trivial way. When existence and continuity theorems are proved locally, we extend them immediately to the
289
Comparison theorem.
entire domain of the right hand side. About the same situation appears when we investigate properties of upper solutions, namely Theorem 5.1. Assume that U is an open convex subset of the product ~ X IR n , F: U - t ~n is a multi-valued mapping satisfying (2.1 - 3), 'Y is an open cover of the set U by sets of the form T x UI X ... X Un, where T, UI , ... , Un are intervals of the real line. Let Z = D{F) E Rce(U) and every set V E 'Y satisfy the condition:
(5.3) if (to, Yo) E V, then there exist c > 0 and a solution Zo of the Cauchy problem y E Zv, y(t o) = Yo on the segment [to, to + c] such that if a function Z = (ZI, ... , zn) E Cs(V) is generalized absolutely continuous, 1f(z) ~ [to, to + c], z(inf1f{z» -< zo(inf1f(z» and z'(t) -< supF(t,z(t» Jor all i = 1, ... ,n and Jor almost all t E 1f(z), then Z -< Zo (on 1f{z) in the sense of the ordering oj §4). Then for every point (tl' yd E U there exists a function that
ZI
E Z+ such
(5.4)
and (5.5) iJ a Junction Z E Cs{U) is generalized absolutely continuous, 1f{z) ~ 1f(ZI), z{inf1f{z» -< zl{inf1f(z)) and z'{t) -< supF{t,z(t» for alli = 1, ... ,n andJor almost allt E 1f{z), then Z -< Zo (on 1f(z»).
Proof. I. Condition (5.3) and the usual procedure for the construction of a maximal extent ion imply the existence of a function ZI E Z+ satisfying the conditions (5.4) and: (5.6) for every point t E 1f{zd there exists neighborhood V E 'Y of the point (t,ZI(t)) satisfying (5.3). II. Fix t2 E 1f(zd. Consider the subset
of the set Ut2 • Denote by W2 the set of all points y E WI satisfying the condition (5.7) if a function Z E Cs{U) is generalized absolutely continuous, 7r(z) ~ 1f{zd, t2 = inf1f{z), Z{t2) = Y and z'(t) -< supF(t,z(t» for almost all t E 1f(z), then there exists c > 0 such that z(t) -< Zl(t), z(t) "# Zl(t) for t E [t2' t2 + c) n 1f(z).
290
CHAPTER 9
By (5.6) there exists a neighborhood G ~ Ut2 of the point ZI (t 2 ) such that G n WI ~ W 2 • Thus W 2 i= 0. Show that WI = W 2. Assume the opposite. By virtue of previous remarks and the connectedness of the set WI this implies the nonemptiness of the set aW2 n WI. Let Y2 be a point of the set aW2 n WI nearest to the point ZI (t2). For the point (t2' Y2) (as (to, Yo)) fix V E'Y and c > 0 according to (5.3). Take a point Y3 E Vt2 of the interval connecting in Ut2 the points Y2 and ZI (t2). Consider a function Zo and a number c > 0, the existence of which is guaranteed by (5.3) for the point (t 2,Y3) E V (as (to,Yo)). Notice that the set W3 = {y : Y E WI n Vt 2, Y -< Y3, Y i= Y3} contains the point Y2· Consider an arbitrary function Z E Cs(U) such that t2 = inf7f(z), z(t 2) E W3 and z'(t) -< supF(t, z(t)) for almost all t E 7f(z). Since z(h) E W3 , there exists a number CI E (O,min{c,sup7f(zo),SUP7f(ZI)}) such that
(5.8) and
(5.9) By (5.8-9) Vt2 ~ W 2. This contradicts the choice of the point Y2 E aw2. Thus our assumption is false and WI = W 2. III. Now take an arbitrary absolutely continuous function Z E Z with 7f(z) ~ 7f(ZI), z(inf 7f(z)) -< ZI (inf 7f(z)) and z' (t) -< sup F( t, z( t)) for almost all t E 7f(z). Denote by G the set {t : t E 7f(z), z(t) -< ZI(t)}. The nonemptiness and the openness of the set G follow from II. As the condition of the ordering which occurs in its definition is non-strong (that is, an equality may be admitted), then the set G is closed. By virtue of the connectedness of the segment 7f(z) this implies that G = 7f(z). This means the fulfilment of (5.5). The theorem is proved. •
CHAPTER 10
CHANGES OF VARIABLES, MORPHISMS AND MAXIMAL EXTENSIONS
In this chapter we look at what aspect the notion of the change of variables takes in framework of our theory. We also take some further steps in the development of our topological tools. 1. Change of variables: general remarks
In a simplest case the change of variables consists in the following. We have a scalar equation y' = J(t,y). We introduce a new variable x related with the old one by a formula
(1.1)
x =
which may be solved with respect to y:
(1.2)
y = 'IjJ(t, x).
Substitution of (1.1) and (1.2) in the initial equation gives a new equation ,
o
x = ot (t, 'IjJ(t, x))
o
+ oy (t, 'IjJ(t, x))J(t, 'IjJ(t, x)),
which becomes the subject of further investigation. As the relation between x and y is known (and is defined by (1.1) and (1. 2) ), results of the investigation of the new equation may then be transferred on the initial one. Of Course, in the realization of this plan we may meet essential technical difficulties, but in general this scheme is applied rather often. A little later on we shall return to such a situation. Now introduce the most general notion which can be understood as a 'change of variables'. Let U be an open subset of the product IR x IRn , Z E R(U). Let
CHAPTER 10
292
to a segment. Therefore cp(z) E Cs(Ud. The homeomorphism 'P is called positive (respectively, negative) admissible change of variables in the space Z if:
(+) the function ~('P(t,z(t))) is increasing (here t E 7r(z) and ~ denotes the projection of the product JR. x JR.n into the first factor JR.), respectively, (-) the function
~('P(t,
z(t))) is decreasing.
So an admissible change of variables (both positive and negative) generates an one to one mapping cP of the set Z onto the subset cp(Z) of the space Cs(Ud. Show that the mapping cP preserves the structure of R(U). 1. If z E Z and a segment I lies in 7r(z), then ZII E Z by the definition of R(U). Therefore the function y = CP(ZII) is defined and Gr(y) ~ Gr(cp(z)). Therefore 7r(y) ~ 7r(cp(z)) and y = cp(z)I,..(y). Since y E Cs(Ud, 7r(y) is a segment. If z E Z and a segment I lies in 7r(cp(z)), then the set Gr(cp(z)II) is homeomorphic to a segment. Hence the set M = 'P- 1 (Gr(cp(z)II)) is also homeomorphic to a segment. Evidently CP(Z)II = cp(zIJ), where J denotes the projection of the set M in JR. (the first factor in the product JR. x JR.n). 2. If ZI, Z2 E Z the domains of the functions ZI and Z2 intersect and the functions coincide on the intersection of the domains, then by the definition of R(U) the function: for t E 7r(zd, for t E 7r(Z2),
7r(Z) = 7r(zd U 7r(Z2), belongs to Z. By the first part of Remark 1 we can point segments 11 ,]2 ~ 7r(cp(z)) such that cp(zd = cp(z)llt and CP(Z2) = CP(Z)II2 • The domains of the functions ZI and Z2 intersect. Therefore the segments II and 12 intersect too. Evidently II U 12 = 7r(cp(z)) and cp(z)(t) =
{~(Zd(t) 'P(Z2)(t)
for t E II = 7r(cp(zd), for t E 12 = 7r(CP(Z2))·
Let ZI, Z2 E Z. Let the domains of the functions cp(zd and CP(Z2) intersect and the functions cp(zd and CP(Z2) coincide on the common part I of its domains of definition. By the second part of Remark 1 there are segments Ij ~ 7r(Zj), j = 1,2, such that CP(Zj)II = CP(ZjIIJ. As cp(zdlI = cp(z2)II and the mapping cp is one to one, then II = 12 and zll I l = z2112 • The sets 7r(cp(zd) \ I and 7r(CP(Z2)) \ I lie in the line on different sides of the segment I. Therefore conditions (+) or (-) imply that the sets 7r(zd \ II
Changes of variables, morphisms and maximal extensions.
293
and 7r(Z2) \ II lie on different sides of the segment II' So 7r(Zl) n 7r(Z2) = II. By the definition of R(U) the function for t E 7r(Zl), for t E 7r(Z2),
7r(Z) = 7r(zr) U 7r(Z2), belongs to Z. We have 7r(cp(z)) = 7r(cp(zd) U 7r(CP(Z2)) and
cp(Z)(t) =
{~(Zr)(t) CP(Z2)( t)
for t E 7r(CP(ZI)), for t E 7r(CP(Z2))'
Our remarks imply, in particular, that cp(Z) E R(Ur). Conditions (+) and (-) were used only in the second part of the second remark. In order to understand their role give the following Example 1.1. Let U = JR x JR, Z be the set of all (defined on segments of the real line) functions of the form z(t) = t + a with non-negative values and of functions of the form z(t) = -t + a with non-positive values (Figure 10.1). The homeomorphism cp of the plane JR x JR onto itself defined by the formula (X,y) for y ~ 0 cp(x,y) = { (x+2y,y) for y ~ 0 leaves functions in the upper half plane fixed and in the lower half plane it 'turns' the functions, transforms them into the functions of the form z(t) = t + a (Figure 10.2). Here cp(Z) (j. R(JR x JR) (naturally, we define the mapping cp following the formula of the definition of an admissible change of variables), because when we extend functions of the lower half plane by functions from the upper half plane we leave the set cp( Z).
Figure 10.1
Figure 10.2
Continue the investigation of properties of admissible changes of variables.
294
CHAPTER 10
Remark 1.1. If Z E Re(U) then 0 according to Lemma 6.6.3 for the one point set {(to, Yo)} as the compactum K. The uniform convergence of a sequence of functions on the segment [to - 8, to +8] is equipotent to the simultaneous uniform convergence of the sequences of their restrictions on the segments [to - 8, to] and [to, to + 8] (see Theorems 3.5.4 and 3.5.5). The condition Z E R(U) implies the homeomorphness of the set M = {z: z E Z, 7r(z) = [to - 8, to + 8], z(t o) = Yo} to the Tychonoff product {zl[to-a,to]: z E M} x {zl[to,toH]: z EM}. Lemma 8.1.1 and Theorem 2.2.3 imply the compactness of the set M. By Theorem 8.1.3 the set M is connected. Since
0 the c:-neighborhood of the point tl in the line ~ does not meet the set B. For every point t E (t l - C:, tl + c:) the set {z(t): z E
Changes of variables, morphisms and maximal extensions.
295
2. Change of variables in equations and inclusions Let us now move to the consideration of a narrower situation outlined in the beginning of the previous section. What happens with equations and inclusions under a change of variables? In the previous section we did not give any convenient description of the transformed space cp( Z). We noticed only that the space Z goes under the change of variables into some space cp( Z). Here the description of the latter space ended. In using a change of variables for an explicit solving of a particular problem we need to have the possibility of continuing the investigation after the change of variables. So we need a good description of the transformed solution space. For instance, it may suit to represent it as a solution space of some (new) equation. We proceed now to the development of corresponding tools. The following assertion is helpful. Lemma 2.1 Let a continuous mapping 9 : V --t IR n , n = 1,2, ... , be defined on an open subset V of the space IRd , d = 1,2, ... , and have continuous partial derivatives. Let a function z : [a, b] --t V, -00 < a < b < 00, be generalized absolutely continuous. Then the function gz : [a, b] --t IR n , (gz)(t) = g(z(t)), is also generalized absolutely continuous, and for almost all t E [a, b] the derivatives z'(t) and (gz)'(t) exist, and
(2.1)
(gz)'(t) =
~! (z(t)) . z'(t).
In the latter formula ~ denotes the Jacobian of the mappings g. The derivatives in t are represented by columns of their coordinates, the multiplication in the right hand side is the matrix multiplication of a matrix by a column. Proof of Lemma 2.1. I. All assertions of the lemma have a local character and it is sufficient prove the lemma for a small part of the domain of the function z. Let to E [a, b]. Find a number c > 0 such that [Oez(to)] ~ V. Find 8> 0 such that Z(025tO) ~ Oez(to) (see Theorem 1.7.3). Put ZI = ZI[to-5,to+5jn[a,bj. By Lemma 6.5.4 the mapping glo.z(to) satisfies the Lipschitz condition. Lemma 4.5.1 implies the generalized absolute continuity of the function gZI· By virtue of the arbitrariness of the point to E [a, b] we obtain the generalized absolute continuity of the function gz. II. Let gl,"" gn be the coordinate functions of the vector function g, andz1 , • .• ,Zd be the coordinate functions of the vector function z. Let the fUnction z have the derivative at a point to with respect to a set M, for which the point to is a density point. Now repeat the calculation which helps to prove an analogous formula in the mathematical analysis (naturally, under
CHAPTER 10
296
our new assumptions). For t E M and i = 1, ... ,n we have
9i(Z(t)) - 9i(Z(tO)) t - to 9i(ZI (t), Z2(t), Z3(t), ... ,Zd(t)) - 9i(ZI (to), Z2(t O), Z3(t O),'" ,Zd(tO)) t - to 9i (Zl (t), Z2 (t), Z3 (t), ... , Zd (t)) - 9i (Zl (to), Z2 (t), Z3 (t), ... , Zd (t)) t - to 9i(ZI(tO), Z2(t), Z3(t), ... , Zd(t)) - 9i(ZI(tO), Z2(t O), Z3(t), ... , Zd(t)) +~~~--~--~----~~~~~~~~~~~--~ t - to 9i(ZI (to), Z2(t O), Z3(t), ... , Zd(t)) - 9i(ZI (to), Z2(t O), Z3(t O),"" Zd(t)) +~~~--~--~----~~~~~~~~--~----~~ t - to O),'" ,Zd-l(tO),Zd(t)) - 9i(ZI(t O),'" ,Zd(tO)) + .. . + 9i(ZI(t ~~~---'----'--'-'------'--'--'--------=----'----'---'-'-------'-~:"""":"':' t - to 89i Zl (t) - Zl (to) = -8 (CI' Z2(t), Z3(t), ... ,Zd(t))--'-'-~---"-.:......::...:. Xl t - to 89i Z2(t) - Z2(t O) + -8 (Zl (to), C2, Z3(t), . .. ,Zd(t)) X2 t - to 89i Z3(t) - Z3(t O) + -8 (Zl (to), Z2(t O), C3, ... , Zd(t) )--,-,-~---,,-.:......::...:. X3 t - to 89i Zd(t) - Zd(t O) + ... + -8 (Zl (to), ... ,Zd-l (to), Cd) , ~
t-~
where Cj, j = 1, ... , d, belongs to the segment with the endpoints Zj(t o) and Zj(t), see Mean Value Theorem 4.10.1 (in this case this is the Lagrange formula). By virtue of the continuity of the partial derivatives of the function 9i and the existence of the derivative of the function Z in the point to with respect to the set M the last expression tends to (2.2)
as t --+ to in M. Thus the derivative of the function 9iZ with respect to the set M at the point to exists and is equal to (2.2). The formula (2.1) is the matrix record of this fact (for all i = 1, ... ,n together). The lemma is proved. • Lemma 2.2. Let -00 < a < b < 00. Let functions Zl, Z2 : [a, b] --+ lR be
generalized absolutely continuous. Then their product ZlZ2 is a generalized absolutely continuous function and for almost all t E [a, b] the derivatives z~(t), z;(t) and (ZlZ2)'(t) exist and (ZlZ2)'(t) = Z~(t)Z2(t) + Zl(t)Z;(t).
Changes of variables, morphisms and maximal extensions.
297
Proof. Define the mapping 9 : ~2 -+ ~ by the formula g(Xl' X2) = XIX2. Its Jacobian has the form of the row (X2' xd. By Lemma 4.5.3 the function z : [a, b] -+ ~2, z(t) = (Zl (t), Z2(t)) is generalized absolutely continuous. The first assertion of Lemma 2.1 implies the generalized absolute continuity of the function (Zl·Z2)(t) = g(z(t)). The second assertion of Lemma 2.1 implies the formula in question (in fact, it is not necessary to use Lemma 2.1 in its full volume and it is sufficient to refer to the equality of the derivative of the scalar function giZ to the expression (2.2), which is established in the proof of Lemma 2.1). The lemma is proved. • Lemma 2.3. Let under the hypotheses of Lemma 2.2 the function Z2 do not vanish. Then the quotient Zl / Z2 is a generalized absolutely continuous function and for almost all t E [a,b] the derivatives z~(t), z~(t) and (ZdZ2)' (t) exist and
( Zl)' (t) = Z~(t)Z2(t)2-ZI(t)Z~(t). Z2 Z2(t) Proof. The proof is analogous to the proof of the previous lemma. Let the function z be defined as in the proof of Lemma 2.2. Let the mapping 9 : ~2 -+ ~ be defined by the formula g(Xl' X2) = XlX2l. In this case the Jacobian of the mapping has the form of the row (X2l, -Xl· X22). Referring to Lemma 2.1 completes the proof. • Lemma 2.4. Let -00 < a < b < 00 and U = (a, b) x ~n. Let functions Aij : (a, b) -+ ~, i = 1, ... , d + 1, j = 1, ... , n, be generalized absolutely continuous. Let
where X = (Xl' ... ' Xd). Let gl, ... ,gn be coordinate functions of the mapping g : U -+ ~n. Then for every generalized absolutely continuous function Z E C.(U) the function 'l/J(t) = g(t, z(t)) is generalized absolutely continuous and for almost all t E 7r(z) all occurring derivatives exist and
+ ... + A~it)Zd(t) + A~d+l (t) + All(t)Z~(t) + ... + Ald(t)Z~(t),
'l/J~ (t) = A~l (t)Zl (t)
(2.3) 'l/J~(t)
= A~I(t)ZI(t) + ... + A~d(t)Zd(t) + A~d+l(t)
+ AnI (t)Z~ (t) + ... + And(t)z~(t), where ZI, ... ,Zd and 'l/JI, ... ,'l/Jn are coordinate functions of the vector functions Z and 'l/J, respectively.
CHAPTER 10
298
Proof. The proof reduces to referring to Lemmas 4.9.2, 2.2 and the remarks of §4.4. • When we write the system (2.3) as one matrix equality, we obtain:
og 'IjJ'(t) = ot (t, z(t))
(2.4)
og
,
+ ox (t, z(t))z (t).
(here ~ denotes the partial derivative of 9 in t, ~ denotes the Jacobian of the mapping gt.) Forget the concrete form of the mapping 9 which is mentioned in Lemma 2.4. As in §1 denote by rp an homeomorphism of an open subset U of the product lR x lRn onto an open subset U1 ~ lR x lRn. Assume in addition that
rp(t, x) = (t, get, x)).
(2.5)
Then the inverse homeomorphism may be represented as
rp-1 (t, x) = (t, h(t, x)).
(2.6)
For every t E lR the mapping gt : Ut - t (U1 )t is an homeomorphism and ht : (U1 )t - t Ut is the inverse homeomorphism. The formula (2.5) implies that rp is a positive admissible change of variables (for the entire space Cs(U) and hence for every space Z E R(U)). Lemma 2.5. Assume that for the change of variables rp from (2.5) and for a generalized absolutely continuous function z E Cs (U) the function 'ljJ = 0(z) ('IjJ(t) = get, z(t)) ) is generalized absolutely continuous. Assume that for almost all t E 7r(z) the derivatives ~(t, z(t)) and ~(t, z(t)) exist and the equality (2.4) holds. Then, if z solves the inclusion
y'
(2.7)
E
then the 'IjJ solves the inclusion , og (2.8) Y (t) E ot (t, h(t, y(t)))
F(t, y), og
+ ox (t, h(t, y(t)) )F(t, h(t, y(t))).
Proof. The proof of the lemma is obvious: (2.8) follows immediately from (2.4) and (2.7) and our notation. • Example 2.1. Keep assumptions made before Lemma 2.5. Assume that the function 9 has continuous derivatives. Let a function z E Cs(U) be generalized absolutely continuous. Let 'ljJ = 0(z). Let Z1, ... ,Zn and 'ljJ1, ... , 'ljJn be the coordinate functions of the vector-functions z and 'IjJ, respectively. By Lemma 2.1 Og1 Og1 Og1 'IjJ~ (t) = ~(t, z(t))z~ (t) + ... + ~(t, z(t))z~(t) + !let, z(t)), VX1
'ljJ~(t)
=
VXn
vt
aa9n (t, z(t))z~ (t) + ... + aa9n (t, z(t))z~(t) + aa9n (t, Z(t)). Xl
Xn
t
Changes of variables, morphisms and maximal extensions.
299
Here gl, ... ,gn denote the coordinate functions of the mapping g. This means the fulfilment of (2.4). Thus under our change of variables an arbitrary solution of the inclusion (2.7) goes into a solution of the inclusion (2.8). Example 2.2. Keep the assumptions of the previous example and require in addition the existence of continuous derivatives of the function h. We now obtain the possibility of repeating the reasoning with respect to the change of variables cp-l. Under this change of variables an arbitrary solution of the inclusion (2.8) goes into a solution of the inclusion (2.7). Thus in this case the change of variables cp transforms the solution space of the inclusion (2.7) onto the solution space of the inclusion (2.8). Example 2.3. Use the notation of the statement of Lemma 2.4. Assume in addition that d = n and the determinant ~(t) =
Anl (t) does not vanish on (a, b). The last condition guarantees an one-valued solvability of the system:
{ with respect to
Xl, ... ,X n :
{ Put
hn(t,y) = Bnl(t)Yl
+ ... + Bnn(t)Yn + Bnn+l(t)
(here Y = (Yl, ... , Yn)). So we have defined the mapping h : U --+ ~n with coordinate functions hl' ... , hn- For every t E (a, b) the mapping h t is the inverse to the mapping 9t· The function .6. is a polynomial of the functions A ij . Lemmas 4.9.2 and 2.2 imply the generalized absolute continuity of .6.. The same lemmas
300
CHAPTER 10
and Lemma 2.3 imply that the functions Bij are generalized absolutely continuous too (by virtue of known formula of Linear Algebra they may be represented as quotients of corresponding determinants, i.e., of polynomials of A j , and ~). Thus the mapping h satisfies all assumptions which first were imposed on the mapping g. By Lemma 2.4 the change of variables
y'(t) = a(t)y(t)
+ a(t)B(t).
All restrictions of Example 8.3.3 are satisfied (here the second term is locally Lebesgue integrable: its measurability follows from Theorem 4.2.6, in order to prove the integrability of this function on an arbitrary segment I ~ (a, b) we can use Lemma 4.7.6 and we can take the function fo(t) = Mla(t)1 as the integrable majorant, where M = sup{IB(t)l: t E I}, see Remark 2.3.5). Now compare our tools with a classical elementary situation. Example 2.6. Consider the system of equations:
The substitution YI = Y2
=
Xl
Xl -
transforms our system into the system (2.9)
+ X2, X2
Changes of variables, morphisms and maximal extensions.
301
The last system 'falls into' independent equations: in the first equation we meet only the function Xl, in the second equation we meet only the function X2. When we solve these equations separately we obtain solutions of the system (2.9). Equations of this type are studied in Example 8.3.3. Apply the result of Example 2.5. We can write the general solution of the system (2.9): Xl
=
cle 5t ,
X2
=
C2e3t,
where Cl and C2 are arbitrary numbers (when we use the formula of Example 8.3.3 we take for the simplicity to = 0). Now we can take to account the made substitution and write the general solution of the initial system:
For the correctness of this change of variables we can here refer to the results of Examples 2.2 or 2.3. Let us return to the consideration of the general situation. Theorem 2.1. Let with the notation of §8.2 F E Qd(
Z~(tl) E
n{ cc( {z~(td: k = j,j
+ 1, ... }):
j = 1,2, ... }.
302
CHAPTER 10
A. Let with the notation of Example 2.3 the derivative ~(t, x) be continuous in x and the elements of the Jacobian ~ do not depend on x. The formula (2.4) goes into the formula
1/J'(t) =
~~ (t, z(t)) + ~! (t)z'(t)
(its concrete type (2.3) is not interesting here). Since the value t = t1 IS fixed, the matrix in the second term is constant. Take an arbitrary E > o. By virtue of the continuity of the function ~(t,x) in x and the convergence zdt 1) ~ zo(td we have
beginning with some k =]1 (see Lemma 1.7.2). In view of the continuity of the linear mapping (with the matrix) ~ (its coordinate functions are linear functionals, therefore they are continuous, see §2.4 and Theorem 2.1.1) there exists a neighborhood 0 z~ (td of the point z~ (td such that II ~ (z~ (t1) - u) II < ~ for u E 0 z~ (td. By virtue of the description of the closed convex hull of a set in §2.4 and by (2.10) for every index] ~]1 there are indices k1 , ... , ks ~ ] and non-negative numbers )'1, ... ,As such that Al + ... + As = 1 and A1ZL (t 1 ) + ... + AszL (td E Oz~(td. Then 111/J~(td
-
(Al1/J~1 (td
+ ... + As1/J~.(td)11
~ Al II ~~ (tl' zo(td) - ~~ (tl' Zk1(td)/I + ... + As /I ~~ (tl' ZO(tl))
-
~~ (tl' Zks (td)/I
+ /I ~! (td(Z~(tl) - (AIZ~1 (t 1 ) + AsZ~s (td))/I E
E
E
< Al-2 + ... + As2- + -2
=
E.
The vector A11/J~1 (t 1 ) + ... + As 1/J~s (tl) belongs to the convex hull of the set {1/JUtd: k = 1,2, ... }. In view of the arbitrariness of the number E > 0 this means the belonging of the vector 1/J~ (td to the set cc( {1/JU t1) : k = 1,2, ... }). Thus cp(Z) E Rn(U). B. Let ~ and ~ be continuous in x. Let a real function ~(t) be such that Ilz~(t)11 ~ ~(t) for every k = 1,2, ... for almost all t E [a, b]. We do not impose on the function ~ any conditions apart from its existence. We need only to state in this way the condition of a corresponding uniform boundedness of derivatives. As the intersection of a countable number of sets of full measure is a set of full measure too, then for almost all
Changes of variables, morphisms and maximal extensions.
303
E [a, bj for every k = 1,2, ... we have Ilz~(t)11 ~ ~(t). Let a point t1 satisfy this condition. Take an arbitrary c > O. By virtue of the continuity of the function ~ in x and the convergence zdtr) -+ zo(tr)
t
beginning with some k = j1' By virtue of the continuity of the linear mapping ~ (t1' Zo (t l )) there exists a neighborhood OZ~(tl) of the point z~(tr) such that for u E OZ~(tl) we have II ~(tl' zo(t))(z~(tr) - u)11 < ~. The continuity of the mapping ~(t, x) in x is in fact the continuity in x of the elements of the matrix of this linear mapping. The coordinate functions of the mapping ~ are linear forms of coordinates of points with elements of the matrix as coefficients. Therefore, the mapping G: UtI x]Rn -+ ]Rn, G(X,U) = ~(tl'X)U, is continuous. Now Theorem 3.2.4, Lemma 1.7.2 and Theorem 3.2.1 imply that on the ball - ~(tr)) the sequence of the mappmgs . a 0/(0, {~(tl' zk(h)): k = 1,2, ... } converges uniformly to the mapping ~ (tl , Zo (tl))' We have
beginning with some k = j2 ~ jl' For every index j ~ j2 there are indices kl' ... ,ks ~ j and non-negative numbers AI, ... , As such that Al + ... + As = 1 and Then 11'¢~(tr) - (AI '¢~I (td
+ ... + As'¢~Jtr))11
~ Alll~~(tl,zo(td) - ~~(tl,zkl(tl))11 + ... +As
11~~(tl,zo(tr)) - ~~(tl,zk.(tl))11
+ II ~! (tl' zo(td)(z~(tr) -
+ Al II (~! (tl' Zkl (tl )) +As c & Al "'" 3
(Al<1 (tr)
~! (tl' zo(t
+ ... + AsZ~. (td))11
l )))
Z~I (tr)11 + ...
11(~!(t1'Zk.(td) - ~!(tl,zo(td)) z~.(tdll c
c
c
c
+ ... + A83- + -3 + Al -3 + ... + A83- = c.
304
CHAPTER 10
The vector
)..1 'ljJ~1 (td
+ ... + )..s'ljJ~. (tl)
belongs to the convex hull of the set {'ljJ~(td: k = ]0,]0
+ I, ... }.
Now as in A we obtain that 0(Z) E Rn(Ud.
3. Maximal extensions. Some generalizations Return to the topic of §6.6. Make here more detailed analysis of relation studied in §6.6. In this section we consider the situation when:
(3.1) U ~ ~ x X is a locally compact subset of the product of the real line ~ and of a metric space X, q : U --t ~ is the restriction to U of the projection of the product into the first factor,
(3.2) Z E
R~(U),
or the following condition reinforcing (3.2):
(3.3) Z E Rc(U). holds. For this situation keep the definition of the symbols Z- and Z+ from §6.6. Assertion 3.1. Let conditions (3.1 - 2) hold and
(3.4)
z E Z.
Then: either (3.5) there exists a function tends the function z;
ZI
E Z nonextendable in the right, which ex-
or (3.6) there exists a function
ZI
E Z+,
which extends the function z.
This assertion is analogous to Lemma 6.6.2. However, we shall give its proof here, which differs a little from the proof of Lemma 6.6.2.
Changes of variables, morphisms and maximal extensions.
305
Proof of Assertion 3.1. Consider the set M = {j: fEZ, n(f) ~ n(z), infn(f) = a, fl 7f (z) = z}, [ where a = inf n(z). Introduce a partial order on the set M and declare that f -< 9 if the function 9 extends the function f, Figure 10.3. a Apply Lemma 1.8.2 to the singleton N = { z }. Let Figure 10.3 Nl ;2 N be a linearly ordered set existing by Lemma 1.8.2 and L = U{ n(f) : fENd. Since functions from Nl extend one other (Figure 10.3), we can define the function Zl : L -+ X by the condition Zll7f(f)
=
f for every function f
E
Nl
.
Evidently inf L = a and if a point s > a belongs to L then [a, s] ~ L. Thus the set L is either a segment [a, b], or a half interval [a, b). I. Consider the case L = [a, b]. By the definition of L there exists a function f E Nl such that b E n(f). Since here n(f) ~ L, n(f) = [a, b]. By the definition of L the function f is nonextendable in the right. Thus (3.5) holds. II. Let L = [a, b). Show that (3.6) is true. The definition of L, Zl and (3.2) imply that zll/ E Z for every segment I ~ L. Show that there exists no function Z2 E Z which extends the function Zl. Assume the opposite. Then the compact urn K = Gr(z2) contains the graphs of all functions from N l . Consider an arbitrary sequence {b i : i = 1,2, ... } ~ L converging to b. Let gi = zll[a,bi]' Since Gr(g;) ~ K for every i = 1,2, ... , condition (c) implies that the sequence {gi : i = 1,2, ... } contains a subsequence {gi: i E A} cona b verging to a function 9 E Z. b1 b 2 By virtue of the continuity of the mapping n we obtain that Figure 10.4 71'(g) = [a, bJ.
b;.-..
CHAPTER 10
306
Since for every i E A we have the coincidence
I
g [a,b;J
=
z21 [a,b;J'
Since both functions g and Z2 are continuous, the last equality implies that g = z21 [a,b]" Therefore bEL, that contradicts our initial assumption (L = [a, b) ~ b). The lemma is proved. • Likewise we obtain: Assertion 3.2. Let conditions (3.1 - 2) and (3.4) hold. Then either
gl[a,b)
=
z21[a,b)"
(3.7) there exists a function Zl E Z nonextendable in the left, which extends the function z, or (3.8) there exists a function Zl E Z-, which extends the function z.
•
Assertion 3.3. Let conditions (3.1-2) hold. Let a function Z : [a, b) ----) X belong to Z+. Then the curve (t, z(t)) goes out from every compact subset of the set U as t ----) b, i. e., for every compactum K S;;; U there exists a number to E [a, b) such that (t, z(t)) E U \ K for all t E (to, b). Proof. The proof of this assertion is completely analogous to the proof of Lemma 6.6.1 and we omit it. • A symmetric assertion is true for functions from Z-. Corollary. Let conditions (3.1 - 2) hold and z E Z- U Z+. Then (3.9) the set Gr(z) is closed in the space U.
•
Assertion 3.4. Let conditions (3.1-2) hold. Let a function z : [a, b) ----) U (respectively, z : (a, b] ----) U ) satisfy condition (3.9) and ZII E Z for every segment I S;;; 7I"(z). Then z E Z+ (respectively, z E Z- ). Proof. We need to check that Z has no function which extends the function z. Consider the case 7I"(z) = [a, b). Assume the opposite. Let a function Zl E Z extend the function z. Then the compactum K = Gr(zl) contains a closed set homeomorphic to the half interval [a, b), which is impossible. The contradiction obtained gives what was required. The case 71"( z) = (a, b] may be considered in an analogous way. Figure 10.5 The assertion is proved. • Denote by Z-+ the set of all functions (a, b) ----) U satisfying conditions (3.9) and
Changes of variables, morphisms and maximal extensions.
307
ZII E Z for every segment 1<;; n(z) (Figure 10.5). The Corollary of Assertion 3.3 and Assertion 3.4 imply immediately: Assertion 3.5. Let conditions (3.1-2) hold. A function z : (a, b) - t X belongs to Z-+ if and only if
zl(a,t] E Z- and Zl[t,b) E Z+ for some t E (a, b).
•
And: Assertion 3.6. Let conditions (3.1- 2) hold. A function z : (a, b) belongs to Z-+ if and only if
-t
U
•
zi(a,t] E Z- and Zl[t,b) E Z+ for every t E (a, b).
Thus the new definition of the symbol Z-+ introduced here is equipotent to the definition from §6.6. Let us continue. Assertion 3.7. Let conditions (3.1 - 2) hold. Let z E Z- (respectively, z E Z+). Then: either
(3.10) there exists a nonextendable in the right (respectively, in the left) function ZI E Z- (respectively, ZI E Z+), which extends the function z; or (3.11) there exists a function
ZI
E Z-+,
which extends the function z.
Proof. The proof of this assertion follows the plan of the proof of Assertion 3.1. Consider the case z E Z-. Denote by M the set { f: f E Z-, supn(J) > supn(z), fl (-oo,SUp1l'(z)]n1l'(f) -- z}
.
As in the proof of Assertion 3.1 we point a maximal linearly ordered subset NI <;; M and we put
L = U{n(J): fENd. Define the function
ZI :
L
-t
U by the formula
zI11l'(f)
=
f
for
f
E
NI
.
H supL E L, then we have the fulfilment of (3.10). This is quite analogous to the situation in the step I of the proof of Assertion 3.1. Consider the case b = supL rt. L. Let a = inf Land c E (a, b).
CHAPTER 10
308
The set A of all limit points of the curve Zl(t) as t . . . . . b lies in q-l(b). This follows immediately from the continuity of the mapping q. If the set A is empty, then the set Gr(zd is closed and Zl E Z-+ by Assertion 3.5. So we have (3.11). Consider the case of a nonempty set A. Let q yEA. By virtue of the local compactness of the set U there are neighborhoods V ~ [V] ~ W of the point y such that the set K = [W] is compact. There exists a point a c b Cl E [c, b) such that z([cl,b)) ~ W, because Figure 10.6 in the opposite case we can repeat the reasoning of the step I of the proof of Lemma 6.6.1. Repeat the reasoning of the step II of the proof of Assertion 3.1. We obtain the existence of a function Xc : [c, b] ......... X extending the function Zl' Then the function for t ~ c, for t ~ c belongs to Z-. Therefore bEL. The contradiction obtained gives what was required. The assertion is proved. • The assertions proved give a pretext to introduce the following notation. Denote: by zne the set of all functions Z E Z satisfying the condition the set Z has no function which extends the function z; by
zne( -)
the set of all functions
Z
E
Z- satisfying the condition
the set Z- has no function which extends the function z; by
zne( +)
the set of all functions
Z
E Z+ satisfying the condition
the set Z+ has no function which extends the function z. As a consequence of Assertions 3.1-2 we obtain Assertion 3.8. Let conditions (3.1 - 2) and (3.4) hold. Then: either there exists a function
Zl
E zne, which extends the function z;
E
zne( -)
or there exist functions
Z;
Zl
and
Z2
E
zne( +),
which extend the function
Changes of variables, morphisms and maximal extensions.
309
or there exists a function
Zl
E
Z-+ which extends the function z.
•
If Z E Ree (U) then the first two cases are impossible. In the connection with the mentioned facts introduce the following conditions:
(r) for every function Z E Z there exists a function extends the function z;
Zl
E
(l) for every function extends the function z;
Zl
E Z- which
Z
E Z there exists a function
(1) for every function Z E Z there exists a function extends the function z.
Zl
E
Z+ which
Z-+ which
4. Convergent space sequences again Let condition (3.1) hold and:
(4.1) Z E Ri(U); (4.2) a
= {Zi:
i
= 1,2, ... } ~ Ri(U).
Recall that following §7.1 we say that the sequence a converges Cs(U) to the space Z, if:
in
(4.3) for every compactum K ~ U and for every sequence Zi E (Zi)K, i E A, there exists a subsequence {Zi: i E A} converging to a function Z E Z.
Assertion 4.1. Let conditions (3.1), (4.1 - 2) hold. Let:
(4.4) the sequence a con verges to the space Z, Zi E Zi, {J = {Zi 1,2, ... }.
i
Then: either 1) for every compactum K ~ U we have Gr(zd ~ U\K, beginning with some i = io,
or
2) there exists a compactum K ~ U such that Gr{zi) ~ K for infinite number of indices i and then the sequence {J contains a subsequence converging to a function Z E Z K; or 3) there are a function Z E Z+ and a subsequence {Zk: k E A} of the sequence {J such that for every point t E 1f(z) we have t E 1f(Zk), beginning with some k = ko, and the sequence {Zk I,,-(zk)n(-oo,t] k E A, k ~ ko}
converges to the function
zl..-(z)n(_oo,t];
CHAPTER 10
310 or
4) there are a function Z E Z- and a subsequence {Zk: k E A} of the sequence (3 such that for every point t E 7f(z) we have t E 7f(zd, beginning with some k = ko, and the sequence {Zk 17r(zk)n[t,oo): k E A, k ~ ko} converges to the function zl 7r (z)n[t,oo); or 5) there are a function Z E Z-+ and a subsequence {Zk: k E A} of the sequence (3 such that for every segment I ~ 7f(z) we have I ~ 7f(Zk), beginning with some k = ko, and the sequence {zkIJ: k E A, k ~ k o} converges to the function
zir
Proof. If the possibility 2) is realized then the assertion follows immediately from (4.3). If 1) is false then the set M = lim top SUPi---> 00 Gr(zi) ~ U is nonempty. In this case we can assume in addition the existence of points Si E 7f(Zi) such that the sequence {(Si' Zi(S;)): i = 1,2, ... } converges to a point m E M (in the opposite case we pass to a corresponding subsequence). Let K ~ U be an arbitrary compactum containing the point m in its interior. For i = 1,2,... let J i be the largest segment satisfying the conditions Si E J i ~ 7f(Zi) and Zi(Ji ) ~ K (condition (c) implies the existence of such a segment). By (4.4) the sequence {zilJi : i = 1,2, ... } contains a subsequence {Zi 1Ji : i E Ad converging to a function Zo E Z. The case when 7f(zo) consists of one point only is trivial (here we have 2) ). In the opposite case the interior of the set 7f(zo) is nonempty. Fix an arbitrary point to E (7f(zo)) n (q(M)). The sequence {Zi : i E Ad contains a subsequence {Zi : i E A 2 } satisfying the additional condition inf 7f(Zi)
< to -
c
< to + c < sup 7f(Zi)
for some c > 0 and for all i E A 2 . Introduce the notation:
Ii =
+
q
zil 7r (z;}n(-oo,tol'
gi = zil 7r (z;}n[to,oo)
s
(Figure 10.7) and consider the following possibilities: A. There exists a compactum K number of the indices i E A 2 •
Figure 10.7
~
U containing Gr(fi) for an infinite
Changes of variables, morphisms and maximal extensions.
311
Evidently here: AI. There exists a subsequence {Ii: i E C 1 } of the sequence {Ii : i E A 2 } converging to a function f with sup 7r(J) = to·
If A is not fulfilled then we repeat reasonings of the proof of Lemma 7.2.1 and we obtain that: A2. There exist a function f E Z-, sup 7r(J) = to, and a subsequence E Cd of the sequence {k i E A 2 } such that t E 7r(Ji) for every point t E 7r(J), beginning with some i = i1 E C1 , and fd[t,to] - t fl[t,to]' Next, if
{h: i
B. There exists a compactum K infinite number of indices i E C1 ,
~
U containing the set Gr(gi) for
then BI. There exists a subsequence {gi: i E C 2 } of the sequence {gi : i E Cd converging to a function 9 with inf7r(g) = to·
If B is not fulfiled, then we repeat reasonings of the proof of Lemma 7.2.1 and we obtain that: B2. There exist a function 9 E Z+, inf7r(g) = to, and a subsequence {gi: i E C 2 } of the sequence {gi : i E Cd such that t E 7r(gi) for every point t E 7r(g), beginning with some i = i1 E C2 , and gil[to,t] - t gl[to,t]' Put for t E 7r(J), z(t) = {f(t), g(t), for t E 7r(g). By Assertion 3.5 and Corollary of Assertion 3.3 z E ZUZ-UZ+UZ-+. Here if z E Z then condition 2) holds, if z E Z- then condition 3) holds, if z E Z+ then condition 4) holds, Ii z E Z-+ then condition 5) holds. The assertion is proved. • Notice also the following three versions of this assertion. Assertion 4.1r. Let conditions (3.1), (4.1- 2) and (4.4) hold, Zi E Zi+' Then: either 1) for every compactum K ~ U we have Gr(zd ~ U\K, beginning with some i = io; or 2) there are a function z E Z+ and a subsequence {Zk: k E A} of the sequence f3 such that for every point t E 7r( z) we have t E 7r( Zk), beginning
312
CHAPTER 10
with some k = ko, and the sequence {Zk I
7r
to the function
( )n(-oo,t ] : E A, Zk
k ? ko} converges
zl7r(z)n(-oo,t];
or
3) there are a function Z E Z-+ and a subsequence {Zk: k E A} of the sequence f3 such that for every segment I ~ 7r( z) we have I ~ 7r( zd, beginning with some k = ko, and the sequence {Zk II: k E A, k ? ko} converges to the function zir • Assertion 4.11. Let conditions (3.1), (4.1-2) and (4.4) hold, Zi E Zi-' Then: either
1) for every compactum K ~ U we have Gr(zi) ~ U\K, beginning with some z = zo; or
2) there are a function Z E Z- and a subsequence {Zk: k E A} of the sequence f3 such that for every point t E 7r( z) we have t E 7r( zd, beginning with some k = ko, and the sequence {zkl 7r (Zk )n [t,oo ): k E A, k ? ko} converges to the function
zl 7r (z)n[t,oo);
or
3) there are a function Z E Z-+ and a subsequence {Zk: k E A} of the sequence f3 such that for every segment I ~ 7r( z) we have I ~ 7r( Zk), beginning with some k = ko, and the sequence {zkI I : k E A, k ? ko} converges to the function zir • Assertion 4.1£. Let conditions (3.1), (4.1-2) and (4.4) hold,
Zi
E Z:+.
Then: either
1) for every compactum K ~ U we have Gr(zd ~ U\K, beginning with some z = zo; or
2) there are a function Z E Z-+ and a subsequence {Zk: k E A} of the sequence f3 such that for every segment I ~ 7r(z) we have I ~ 7r(zd, beginning with some k = ko, and the sequence {Zk II: k E A, k ? ko} converges to the function zir • As a consequence of Assertion 4.1 we obtain: Theorem 4.1. Let conditions (3.1), (4.1 - 2) and (4.4) hold and a ~ R!(U). Then Z E R!(U). • The corresponding one sided versions of Theorem 4.1 are true too.
Changes of variables, morphisms and maximal extensions.
313
5. Morphisms
Let us now expand the notion of a change of variables and consider the notion of a morphism corresponding to the circle of questions under consideration. Let (5.1) each of the triples ql : U1
----t
~
and q2 : U2 ----t
~
satisfy (3.1).
Generalize the notion introduced in §1. We say that the mapping
(+) the function q2((z). Remark 5.1. Let condition (5.1) hold, Z E Ri(Ud, and a continuous mapping : U1 ----t U2 be extendable to the space Z. Then (Z) E Ri(U2 ). This is obvious. A mapping : U1 ----t U2 is called perfect, if: (5.2) the set (M) is closed in U2 for every closed subset M of the space U1 ; (5.3) the set -l(M) is compact for every compact subset M of the space U2 • Assertion 5.1. Let condition (5.1) hold,
(5.4) Z E R~(Ud, (5.5) a perfect mapping : U1
----t
U2 be extendable to the space Z.
Then (Z+) = ((Z))+, (Z-) = (~(Z))- and ~(Z-+) = (~(Z))-+. _ Proof. The inclusions ~(Z-) ~ (~(Z))-, ~(Z+) ~ (~(Z))+ and ~(Z-+) ~ (~(Z))-+ follow from Corollary of Assertion 3.3 and from (5.2).
CHAPTER 10
314
Show that ((Z))+ c (Z+). Take an arbitrary z E ((Z))+. Let 1f(z) = [a, b). Fix an arbitrary sequence of points t1 < t2 < ... of the half interval [a, b), ti ----t b. Since zl[a,t;j E ((Z))+ there exists a function Yi E Z such that
(Yi) = zl[a,ti]' Apply Assertion 4.1 with Zi i = 1,2,00.}.
== Z to the sequence of functions {Yi :
By (5.3) the set -1 (Gr (z![a,tlJ)) is compact. Therefore the case 1) of Assertion 4.1 does not take place. The case 2) of Assertion 4.1 does not take place because the compact urn (K) there contains a closed noncompact set Gr(z). The cases 4) and 5) are impossible because for every i = 1,2'00' the point (inf(1f(Yi)), Yi(inf 1f(Yi))) belongs to the compactum -1 (inf 1f(z)), z(inf 1f(z))) which implies the fulfilment of an analogous condition for the limit function. Only case 3) remains. Let Y E Z+ be a corresponding limit function. For it (inf 1f ( z ), Y (inf 1f ( z ))) E -1 (inf 1f ( z ), z (inf 1f ( z ) )) . Take an arbitrary point t E [a, b). Since t ~
=
< ti,
beginning with some
~o,
(Gr(Yi)) 2 Gr (zl[a,tJ)
for
i
~
io·
The last observation and the compactness of the set -1 (Gr (zl[a,tJ)) imply that (Gr(y)) 2 Gr (zl[a,tJ) . In view of the arbitrariness in the choice of the point t E [a, b) this means that (Gr(y)) 2 Gr(z). If for some s E 1f(zr} we have (s, Zl(S)) rt Gr(z), where Zl = (y), then the closed noncompact set Gr(z) turns out to be a subset of the compactum Gr (zl7r(z')n( -00,8J)' which is impossible. Thus (Gr(y)) = Gr(z). Therefore z = (y). Likewise we show that (Z-) 2 ((Z))- and (Z-+) The assertion is proved. The following assertion is obvious. _ Assertion 5.2 Let conditions (5.1) and (5.4 - 5) (Z) E R~(U2)' As a direct consequence of Assertions 5.1-2 we obtain: Theorem 5.1. Let condition (5.1) hold. Let Z E R~f(Ur}, Z E R~r(Ul))' Let condition (5.5) hold. Then (Z) E R~f(U2) Z E R~r(U2).) Theorem 5.2. Let condition (5.1) hold. Assume that:
=> ((Z))-+. •
hold.
Then •
(respectively, (respectively, •
Changes of variables, morphisms and maximal extensions.
(5.6) for i = 1,2, ... a continuous mapping
-7
315
U2 is extendable to
(5.7) a sequence of spaces {Zi: i = 1,2, ... } converges in U l to a space Z;
(5.S) U is an open subset of U2 , a mapping
-7
U is extendable to Z;
(5.9) for every compactum K ~ U l the sequence of the mappings
(5.10) for every compactum K
~
U
a) the set
U{
+ I, ... } ~ W.
Then the sequence {(
(5.11)
Gr((j) ~ W for j E A, j ? i.
Conditions (5.7) and (5.11) and the compactness of the set [W] imply that the sequence {(j : j E A, j ? i} contains a subsequence {(j : j E Ad converging to a function ( E Z[W]. II. Apply Theorem 2.4 to the sequence of the mappings {
j E A 2 } to the function z.
III. In view of the arbitrariness in the choice of the compactum K ~ U and of the sequence {Zj : j E A 2 } II implies the convergence of the sequence
316
CHAPTER 10
of the spaces {(CPi(Zi))U: i = 1,2, ... } to the space (cp(Z))u. The theorem is proved. • Let us highlight the following particular case of Theorem 2.1. Corollary. Let conditions (5.1), (5.7) hold. Let: an homeomorphism CPi : U1 --t U2 be a change of variables in the space E Ri(Ud, for i = 1,2, ... ;
Zi
an homeomorphism cP : U1
Z
E
--t
U2 be a change of variables m the space
Ri(U1 ).
Let condition (5.9) hold and: for every compactum K ~ U2 the sequence of the mappings cp;llK converge uniformly to the mapping cp-1IK. Then the sequence of the spaces {CPi(Zi): i = 1,2, ... } converges in the space R i (U2 ) to the space cp(Z). •
6. One more remark about the continuity of the dependence of solutions on the right hand side The passage to general structures expand our possibilities of investigating singularities. In this section we discuss one such situation. Lemma 6.1. With the notation of (3.1) let V be an open subset of the set U, Zi E Zi for i = 1,2, .... Let {Z} U {Zi: i = 1,2, ... } ~ R(U). Let the sequence {Zi: i = 1,2, ... } converge to a function Z E Cs(U). Let the sequence of the spaces {(ZJv: i = 1,2, ... } converge in V to the space Zv. Then if a segment I lies in 7f(z) and Gr (ZII) ~ V we have ZII E Z. Proof. By Lemma 2.4.2 there exists a neighborhood Va of the set Gr (zIJ, the closure of which is compact and lies in V. Let h = 7f(zd n I if 7f(zd n I f- 0, and h = {t: t E 7f(Zk), p(t,I) = P(7f(Zk),I)} if 7f(Zk) n I = 0. In the second case the segment h degenerates into an one point set consisting of one of the endpoints of the segment 7f(Zk) (nearest to 1). Since by Theorem 3.5.1 7f(zd --t 7f(z) (that is inf 7f(Zk) --t inf 7f(z) and sup7f(zd --t SUp7f(z)), h --t I. By Theorem 3.5.4 zkl h --t z/r
IJ
By the Corollary of Theorem 3.4.2 Gr ( Zk / ~ Va, beginning with some k = k o. Thus Zk /h E Cs([Vo]). The definition of the convergence of a space sequence implies now that the limit function Z/I belongs to Z(Vo]· The lemma is proved. •
Changes of variables, morphisms and maximal extensions. Let Z ~ Cs(U). A continuous mapping
--+
317
U is called a
(6.1) the mapping (Z) ~
z.
Conditions (6.1-3) imply, in particular, that {Z) = Z
(6.4) F: U
--+ ~n
(6.5)
--+
be a multi-valued mapping;
U be a D(F)-retraction;
(6.6) {Zi: i = 1,2, ... } E s{U); (6.7) for i = 1,2, ... a continuous mapping
--+
U be extendable to
(6.8) for every compactum K ~ U the sequence {
i = 1,2, ... }
(6.9) the sequence {(Zi)U\
(6.10) the sequence {i{Z;): i = 1,2, ... } converge in
Proof. 1. Notice first that (6.8) implies that for every compact K ~ U the closure of the set U{
318
CHAPTER 10
IV. Apply Theorem 3.2.4 to the sequence of mappings {i; : j E A 2 } and to the sequence of compacta {Gr( Zj) : j E A 2 }. Condition (6.8) implies that the sequence of the partial mappings {i; Icr(z;) : j E A 2 } converges to the partial mapping lcr(z)" This implies the convergence of the sequence of the compacta {dGr(Zj»: j E Ad to the compactum (Gr(z». Therefore the sequence of the functions {(U) is closed in U (see Lemma 1.7.4 and (6.1», therefore the set M = {t: t E 7r(z), (t,z(t» E (U)} is closed in 7r(z). Condition (6.9) and Lemma 6.1 imply, that for every segment I ~ 7r(z) \ M we have ZII E D(F). Therefore z'(t) E F(t, z(t» for almost all t E 7r(z) \ M. Since Zo E D(F) and Zo 1M = zI M , z'(t) = Z~ E F(t, zo(t» = F(t, z(t» for almost all t E M (recall, that we mean here approximate derivatives). That is, z'(t) = F(t, z(t» for almost all t E 7r(z). Therefore Z E D(F). The assertion is proved. • The assertion proved implies: Theorem 6.1. Let conditions (6.4 - 6), (6.9) hold. Let:
(6.12) for i = 1,2, ... a continuous mapping i : U the space Zi;
---?
U be extendable to
(6.13) the sequence {(U) to the space (D(F» = D(F, (U». Then condition (6.11) holds. Theorem 6.2. Let condition (6.4) hold. Assume that:
•
(6.14) Zo E (D(F»-+ (that is, Zo is a maximaly extended solution of the inclusion y' E F(t, y») and all limit points of the graph of the function Zo as the argument tends to the endpoints of 7r(zo) lie outside U); (6.15)
{Zi: i = 1,2, ... }
~
Ri(U);
(6.16) the sequence {(Zi)U\Cr(zo): i the space D(F, U \ Gr(zo».
= 1,2, ... }
converges in U \ Gr(zo) to
Then condition (6.11) holds. Proof. In order to prove (6.11) it is sufficient to point an open cover of the set U, for every element V of which (Zi)V ---? D(F, V), see Theorem 7.6.4. Therefore it is sufficient to prove the convergence (Zi)Uo ---? D(F, Uo), where Uo = un (7r(zo) X ]Rn).
Changes of variables, morphisms and maximal extensions.
319
For (t, y) E Uo put cI>(t, y) = (t, zo(t)). Let cI>i = cI> for all i = 1,2, .... The fulfilment of (6.6) follows immediately from the Corollary of Theorem 7.2.1. The fulfilment of (6.7-9) is trivial. Condition (6.10) holds because ~i(Zi) ~ D(F, Gr(zo)) = ~(D(F)). The theorem is proved. • Example 6.1. Let a scalar function f be defined on the plane JR x JR, be continuous and f(t, 0) = 0 for all t E lR. Then solutions of the equation y'
= f(t, y) + a
2: 2 y
depend continuously on the parameter a. For a i- 0 this follows from classical theorems on the continuity of the dependence of solutions on parameters (see Theorem 7.1.3). To prove the assertion for a = 0 consider an arbitrary sequence ai --+ O. Let Zi denote the solution space of the equation y
,
ai = f (t, ) Y + 2 2' a i +y
Put U = JR x JR and Zo = O. The fulfilment of (6.4) and (6.14-15) is obvious. The fulfilment of (6.16) follows from classical continuity theorems. Thus we obtain what was required. As a direct consequence of Theorem 6.1 we obtain: Theorem 6.3. Let conditions (6.4 - 5) hold, D(F) E Rs(U),
(6.17)
D(F, U \ (U))
E
Rc(U \ cI>(U)).
Then
•
(6.18) D(F) E Rc(U).
As a consequence of this theorem we obtain Theorem 6.4. Let conditions (6.4) and (6.14-15) hold. Then condition (6.18) holds too. • Example 6.2. Let a scalar function f be defined on the plane JR x JR, be continuous on the set JR x (JR \ {O}) and f(t,O) = 0 for all t E JR. Then solutions of the equation y' = f(t, y) depend continuously on initial values. This may be established with the help of a direct reference to Theorem 6.4. We may repeat the arguments of Example 6.1 and establish that solutions of the equation y' = f(t, y)
+a
2: 2 y
(with the new assumptions about f) depend continuously on the initial values and on the parameter a.
320
CHAPTER 10
Example 6.3. Consider the two-dimensional system:
(6.19)
{
xI =
I 1 (t, ) a 2 sin2 y x + ------.::~
y'
a 2y2 12(t,X,y) + a 6 + (2 y-a 2)2'
=
a4
+ t2 + X2,
where the functions 11 and h are define and continuous for all real value of t, x and y. Let 12(t, x, 0) = 0 for all t, x E lR. For a = 0 we put the fraction identically equal to zero. Pay attention to a particularity of the dependence of the right hand sides of the system (6.19) on the parameter a: the second terms of the equations have discontinuities of the type of passage to the infinity (as in previous examples). But the dependence of the solutions on the parameter a is continuous. For a i= 0 this follows from classical theorems. To prove the continuity for a = 0 consider an arbitrary sequence ai --t O. Let Zi denote the solution space of the system 2
(6.20)
{
• 2
s lll y x I = I 1 (t,x ) + __a.....:ic..._ ---.:,-a 4., + t 2 + X2, a 2y2 y' = 12(t, x, y) + --::-6-("'-:":'2~--::-:2)-2'
ai
+
y - ai
Define the mapping cI> : ]R X ]R2 --t ]R X ]R2 by cI>(t, x, y) = (t, x, 0). Put cI>i == cI> for i = 1,2, .... Here the space (Zi) turns out to be a subset of the space Ii of solutions of the differential inclusion
(6.21)
x' E11(t,X) + [0, ai4+a!t + x 2] .
The sequence of solution spaces {Ii : i = 1,2, ... } converges in the plane of the variables (t, x) to the solution space of the equation
(6.22) as i --t 00. We have here the simplest case of the one point singularity. Thus (6.10) holds. The fulfilment of (6.4-5) and (6.7-9) is obvious. To obtain the possibility ofreferring to Assertion 6.1 it remains to prove the fulfilment of (6.6). But it is obvious, because outside the line x = 0 the corresponding convergence follows from classical theorems and in a small neighborhood of an arbitrary point of the form (8,0) every solution of the system (6.20) is also a solution of the system consisting of the equation (6.21) and of the inclusion
(6.23)
Changes of variables, morphisms and maximal extensions.
321
In order to prove the convergence of the sequence of solution spaces of the equations (6.23) as i - t (X) to the solution space of the differential inclusion y' E [-10,10] we can use Theorem 6.2, because the function z == 0 satisfies the inclusion y' E [-10,10]. This fact and an analogous convergence of the sequence of the solution spaces of the equations (6.21) to the solution space of the equation (6.22) imply the convergence of the sequence of the solution spaces of the systems of the equations (6.21), (6.23) to the solution space of the system
{
(6.24)
X' = fl(t,X), y' E [-10,10]
(in a small neighborhood of the point (t, x, y)). Then (independently on concrete type of the system (6.24) and only as a consequence of the fact of the convergence to a solution space) we obtain the fulfilment condition (6.6). So the continuity of the dependence of solutions of the system (6.19) on the parameter a is established. 7. Uniqueness theorem
We give here, perhaps a somewhat unexpected application of ideas of this chapter. Consider a more general situation. Let:
(7.1) Xo be a Hausdorff topological space; (7.2) X =
~
x X o, q : X
-t
~
be the projection on the product into the
first factor;
(7.3) Yo be a locally compact metric space and Y = ~ x Yo;
(7.4) Yo E Yo and p : Y factor; (7.5) Z E Ri(X),
-t
~ be the projection of the product into the first
Zo E Z,
0 E (7r(zo));
(7.6) {Uj : j = 1,2, ... } be a base of neighborhoods of the set Gr(zo) in the space X n q-l(7r(ZO))'
CHAPTER 10
322
(**) for every neighborhood 0Yo of the point Yo in the space Y we have 'Pj(aUj ) ~ lR. x (Yo \ 0Yo), beginning with some j, and
(7.7) 'Y be a family of open subsets of the space Y, moreover every element V of the family 'Y do not contain the point (0, Yo) and satisfy the condition
(7.8) the set H contain the set limtopsup{Gr(z): Z E
~
(-E,E), z(O) = Yo}.
€-+O,j-+oo
Theorem 7.1. Let conditions (7.1 - 8) hold. Let the connected component of the point (0, Yo) in the set (H \ U'Y)o be trivial (i.e., consists of one point (0, Yo) only), Z E Z, 0 E 7r(z) and z(O) = zo(O). Then zi [-t: ,t: ]n7r(Z )n7r(ZO) = zol [-t: ,t:]n7r(Z )n7r( zo) for some E > O. Proof. Assume that it is false. Then we can fix points ti E 7r(z) such that ti -+ 0 and Ilz(tdll =1= O. The continuity of the function z implies that IIz(ti)1I -+ O. It is sufficient to consider the cases when either all points of the sequence {ti: i = 1,2, ... } lie on the left of 0, or all such points lie on the right of O. Restrict consideration to the first case. By (7.6) we can assume in addition that (ti' z(t i )) E aUi and that Gr (zl(t"o]) ~ Ui · Fix an arbitrary neighborhood 0Yo of the point Yo with the compact closure. By (7.6,**) 'Pi(ti,Z(t i )) tt. lR. x 0Yo, beginning with some i = io. Then we can fix points Si E [ti' 0] such that Zi(Si) E aOyo and Zi((Si, 0]) ~ 0Yo for Zi =
Changes of variables, morphisms and maximal extensions.
323
Theorem 7.1, as it is stated, is ready for direct use in proofs of the uniqueness of a solution of the Cauchy problem. Here we need to take as X the domain of definition of the right hand side of the equation under consideration. As Y we may take the space JR.k (often for k = 1, see the discussion of this question in [HaJ). Often it is sufficient to take as the set H the entire space Y. In this case (7.8) holds. In order to show better the situation let us give the statement of two particular cases of Theorem 7.1 and examples helping to understand their use. In the discussion of the question about the uniqueness of a solution of the Cauchy problem a particular case of the uniqueness of the zero solution turns out to be in the center of attention, see [HaJ. According to this remark let U denote below an open subset of the product JR. x JR.n, Y = JR. k , Yo = 0, Zo == O. Theorem 7.2. Let Zo E Z E Ri(U). Let conditions (7.6 - 8) hold. Let is the connected component of the point (0, Yo) in the set ({O} x JR.n) \ trivial. Let z E Z, 0 E 7l'(z) and z(O) = O. Then zl[ -E,£n7r€ J () == 0 for some c> O. • The most complicated moment in the application of Theorems 7.1 and 7.2 may be the pointing of the family,. Notice in this connection that above we have shown possibilities of the verification of the fulfilment of condition (7.7*). In particular, Theorem 7.2 implies: Theorem 7.3. Let Zo E Z E R~(U). Let condition (7.6) hold. Let rpi(Ui ) = U for every i = 1,2, ... , the sequence {CPi(Z): i = 1,2, ... } be convergent (in the space R~(U)), z E Z, 0 E 7l'(z), and z(O) = O. Then zl[ -e,c Jn1i () == 0 for some c > O. • Z Let us show the relation of the results obtained to the classical uniqueness theorem with the Lipschitz condition. Notice first that the differential inequality (inclusion)
U,
Z
(7.9)
Ily'(t)11
~
Llly(t)11
locally satisfies the Davy conditions. Therefore the space Z of its solutions belongs to Rce(JR. x JR.n), see §6.4. The inequality (7.9) goes into itself under the change of variables rp>.(t, y) = (t, >..y) (where>.. E JR.). Thus {rp,\(Z) : >.. = 1,2, ... } --t Z in the space Rce(JR. x JR.n). This implies the fulfilment of the hypotheses of Theorem 7.3. Therefore the solution of the Cauchy problem for the inequality (7.9) with zero initial values is unique (and is identically equal to 0). If now the right hand side of the equation (7.10)
y'
= f(t, y)
satisfies the Lipschitz condition in y with a constant L, Yo (t) and Yi (t) are two solutions of (7.10) on a segment [a, bj, then the function
324
CHAPTER 10
z(t) Yl (t) - yo(t) is defined on the segment [a, b] and satisfies (7.9). Now the previous remark implies the uniqueness theorem for the equation (7.10). Let us now show with simple examples how we may apply the obtained assertions directly to particular equations. Example 7.1. Consider the scalar equation
(let its right hand side be equal to zero for y = 0). In order to prove the uniqueness of a solution of the Cauchy problem with zero initial values consider the mappings CPk(t, y) = (t, ye 7rk ). Notice that under this change of variables the equation under consideration goes into itself. We can take as 'Y the family of all bounded open subsets whose closures do not meet the lines y = ±e7rk , k = 0, ±1, ±2, .... Example 7.2. Let c > O. Consider the scalar differential inclusion y' E F(y), where
for y t/: {±e 7ri , i = 0, ±1, ±2, ... } U {O} and F(y) = ~ in the opposite case. Here the situation is quite analogous to that of the previous example. Example 7.3. Let a function g(t, y) be continuous and bounded. Consider the scalar equation
,
Y
=
yg(t, y) sin2 (ln Iyl)
(let its right hand side be equal to zero for y = 0). Here we obtain the uniqueness of a solution of the Cauchy problem with zero initial from the remark of the previous example, because the solution space of the equation under consideration (locally) lies in the solution space of the inclusion of Example 7.2. Example 7.4. Consider the scalar equation y' = yltlsin(ln Iyl). By arguments analogous to those of Example 7.1 the Cauchy problem with zero initial values has a unique solution. Notice that this equation is not covered by the Caratheodory's theorem. For our theory singularities of its right hand side are not essential, see Example 8.6.l. In Examples 7.1-4 the right hand side of the equation has a discontinuity. Let us modify it in order to keep its idea and to pass with respect to exterior criteria into the domain of the classical theory, but in such a way that assertions of the classical theory be not applicable.
Changes of variables, morphisms and maximal extensions.
325
Example 7.5. Consider the scalar equation y' -
y
- v1YI + sin2 (ln Iyl)'
The change of variables of Example 7.1 transforms our equation into the equation
Y
I
Y = A-!
v1YI + sin
2
(ln Iyl)'
The solution space of this latter equation converges as A = errk solution space of the equation I
Y =
2
Y
sin (In Iyl)
- t 00
to the
•
When we use Theorem 7.3 we obtain the uniqueness of the solution of the Cauchy problem with zero initial values. Example 7.6. Consider the system {
(7.11)
X'
= X
y'
=
.yX"e-lyl/x .
The change of variables (t, y) system X'
(7.12)
{
=
X
+ y + ~e-IYI/X2 -t
(t, AY) transforms this system into the
+ y + Ai .yX"e-Alyl/x2
y' = Ai .yX"e-Alyl/x2.
Let Ak - t 00 and Zk denote the solution space of the system (7.12) with A = Ak' Let U = ~ X (~2 \ {O}). Show that the sequence a = {(Zdu: k = 1,2, ... } belongs to s(U). Outside the x-axis this trivial (the right hand side is uniformly bounded on compact subsets). In a small neighborhood V of a point (s,O), s > 0, of the x-axis the absolute values of x and yare bounded from above by a constant m> O. Define the continuous mapping h : V - t ~ by the formula h(x, y) = x - y. Then if u = (x, y) is a solution of the system (7.12) with values in V and v(t) = h(u(t)) then v' = x' - y' = X +y, and therefore Iv/l :s;: 2m. When we apply Assertion 7.2.1 and Theorem 7.2.1 we obtain the membership of the sequence a in s(U). When we now apply Theorem 7.2 we obtain the uniqueness of a solution with zero initial values. The right hand sides in the system (7.11) are continuous. The example shows no possibility of the investigation of an equation with a discontinuous
326
CHAPTER 10
right hand side, but draws our attention to the possibility of a 'classical' situation where classical results 'do not work' but our methods are applicable. Notice the particularity of the system (7.11) at the point (t, 0,0). When we take y = 0 we obtain that, for instance, in the second equation the right hand side goes into fIX. Write a scalar equation with an analogous right hand side: x' = fIX. For this the uniqueness theorem is not true: besides the solution x == 0 the function
is also a solution of the Cauchy problem with zero initial values.
CHAPTER 11
SOME METHODS OF INVESTIGATION OF EQUATIONS
In this chapter we consider some of the simplest cases of the application of developed tools. We also compare our methods with the classical theory.
1. Linear equations (systems) We speak here about a 'normal' system of linear equations: y~ = a11 Yl
(1.1)
{
+ a12Y2 + ... + alnYn + aI,
y~. ~.~~~~~ .~.~~~~~ .~.·.·.·.~.~~~~.n. ~ ~2' y~ = anI Yl
+ a n 2Y2 + ... + annYn + an,
where real or complex functions aij and ai, i, j = 1,2, ... , n, are defined on a interval (a, b) (a < b) of the real line; moreover, the functions aij are (locally) Lebesgue integrable and the functions ai are Denjoy integrable. The system (1.1) is equivalent to the vector linear equation (1.2)
Y'
= Ay
+ a,
where
y~ (~J C'
a21
A=
.
anI
a12
a22
cr," ) a2n .
a n2
ann
,
a~ (jJ
Both in the real and in the complex cases (see §6.1) we can apply our theory to the equation (1.2). The function n
cp(t) = n
L
laij(t)1
i,j=l
is (locally) Lebesgue integrable on (a, b). The function A(t)y locally satisfies the Caratheodory conditions: for every positive number m the estimate IIA(t)YII ~ mcp(t) is true in the region lIyll ~ m. Thus on the set Urn =
328
CHAPTER 11
{(t,y): a < t < b, Ilyll < m} the right hand side of the equation (1.2) satisfies the hypotheses of Theorem 10.2.1. This implies the fulfilment of the condition ZUm E Rcekn(Urn) for the solution space Z of the equation (1.2). By remarks of §8.3 Z E Rcekn(U), where U = (a, b) x]Rn (respectively, U = (a, b) x en). For a == 0 the equation (1.2) (respectively, the system (1.1)) is called homogeneous. By virtue of the estimate IIAy* - AY**II = IIA(y* -
Y**)II : :;
-
Y**II
we can apply Theorem 6.5.1 to the (homogeneous) equation
(1.3)
y' = Ay
on the set Urn. By Theorem 6.5.1 and Lemmas 6.5.1 and 6.5.3 the solution space of the equation (1.3) satisfies condition (u). This result may be easily expanded on 'non-homogeneous' equation (1.2). Here we can use: Lemma 1.1. Let functions Zl and Z2 solve the equation (1.2) on a segment I. Then the function Z = Zl - Z2 (defined on the same segment 1) solves the equation (1.3). Proof. The proof reduces to direct verification. By Lemma 4.9.2:
z'(t) = z~(t) - z;(t) = A(t)Zl(t)
+ a(t)
- A(t)Z2(t) - a(t)
= A(t)Zl(t) - A(t)Z2(t) = A(t)(Zl(t) - Z2(t)) = A(t)z(t)
(for almost all t E 1), which was required. • With the notation of Lemma 1.1 now let to E I and Zl (to) = Z2 (to). Then z(t o) = O. The function z* == 0 solves (1.3). We have already shown that the solution space of the equation (1.3) satisfies condition (u). Therefore Z == 0, Zl = Z2. This means that the solution space of the equation (1.2) satisfies condition (u). The last observation and Theorem 10.2.1 imply: Theorem 1.1. The space Z of solutions of the equation (1.2) satisfies conditions (c), (e), (u) and (n). • Lemma 1.2. Let non-negative functions u and v be defined on a segment [p, q] of the real line. Let the functions u and uv be Lebesgue integrable, c > 0 and for almost all t E [p, q]
u(t),; c+
!
v(s)u(s)ds (respectiVelY, v(t)';
Then for almost all t E [p, q] •
J u(s)ds
vet) :::;; ce p
c+
i
V(S)U(S)dS) .
J u(s)ds ( respectively, vet) :::;; ce'
q
)
.
329
Some methods of investigation of equations.
If the function v is continuous then these inequalities hold for all t E [p, q]. Proof. For Vet) = c + J; v(s)u(s)ds we have vet) ::::; Vet) and V'(t) = v(t)u(t) ::::; V(t)u(t) for almost all t E [p, q]. Since here Vet) ~ c> 0, V'(t)
(In Vet))' = Vet) ::::; u(t), t
In Vet)
t
= In V(p) + j(ln V(s))'ds ::::;
In V(p)
+j
p
u(s)ds,
p
vet) ::::; Vet) ::::; V(p)eJ>(S)dS = ceJ>(S)dS. The second estimate may be obtained in an analogous way. In the case of the continuity of the function v we have the closedness of the set of all t E [p, q], in which our inequalities are fulfilled. This easily implies the last assertion of the lemma. The lemma is proved. • Theorem 1.2. Let Z denote the solution space of the equation (1.2). All elements of the space Z-+ are defined on the entire interval (a, b). Proof. 1. Consider the particular case of an homogeneous equation. Let z E Z-+ and ?fez) = (aI' bd· Assume that (aI, bd # (a, b). Let for definiteness bl < b. Take an arbitrary point p E (aI' bd. Apply Lemma 1.2 with c = IIz(p)11 + 1, u(t) = cp(t) and vet) = IIz(t)lI. Since for almost all t E [p, bd
IIz'(t)1I ::::; cp(t)IIz(t)II, IIz(t) II = IIz(p) + (z(t) - z(p)) II ::::; IIz(p)1I + IIz(t) - z(p)1I t
=
j z'(s)ds
+
IIz(p)1I
p
t
+
::::; IIz(p)II
j cp(s)IIz(s)IIds p
t
< c+
J
cp(s)lIz(s)lIds,
p
II Z (t)ll ./"":: ce l'
lIence
""::
b1
p
CHAPTER 11
330
which contradicts Lemma 6.6.1. The case a < a1 may be considered in an analogous way using the second version of Lemma 1.2. Thus our assumption is false. This gives what was required. II. Consider the case of a non-homogeneous equation but with an additional assumption that the function CY is (locally) Lebesgue integrable. It is more convenient to work not with the equation (1.2), but with the equivalent system (1.1). Let Z = {Zl,"" zn} E Z-+. Then z* = {Zl' ... ,Zn, Zn+l} E Z;-+, where Zn+1 == 1 and Zl denotes the solution space of the system y~ = CYllY1
Y;
= CY21Y1
+ CY12Y2 + ... + CY1nYn + CY1Yn+1, + CY22Y2 + ... + CY2nYn + CY2Yn+1,
By I the domain of z* coincides with the interval (a, b), that gives what was required. III. Finally, consider the general case. Let cy* be an arbitrary primitive of the function CY. Under the change Y = x + cy* (see Example 10.2.4) the equation (1.2) goes into the equation x' = Ax + Acy*. We obtain what was required from II and remarks of Example 10.2.4. The theorem is proved .• Theorem 1.3. Let Z denote the solution space of the equation (1.2). Let Zl denote the solution space of the equation (1.3). Let Zo E Z-+. Then Z-+ = Zo + Z;-+. The set Z;-+ is a linear space (with respect to usual operations of the addition of functions and of the multiplication of functions by numbers). Proof. The proof is completely analogous to the proof of Lemma 1.1.. Evidently an analogous assertion is true for sets of solutions with a fixed segment I as the domain of definition. But here it is necessary to notice that by Lemma 6.6.2 and Theorems 1.1-2 Z n 7[-1(1) = {ZII: Z E Z-+}. Thus the analog of Theorem 1.3 for a segment is a direct consequence of this theorem. The mapping cI> I : Z;-+ ---. Zl n 7[-1 (1) which associates to a function Z E Z;-+ its restriction zlI to the segment I is linear. The previous remark and condition (u) imply that it is an isomorphism. This is also true for the particular case where the segment I consists of one point c only. The structure of the space Z;-+ n 7[-1 (c) is obvious: it is isomorphic to the space ]Rn (In the complex case it is isomorphic to the space This isomorphism associates to a function Z the vector z(c)). So: Theorem 1.4. With the notation of Theorem 1.3 the dimension of the space ZI+ is equal to n. If el, ... ,en is an arbitrary basis of the space lRn
en.
Some methods of investigation of equations. (respectively, en), c E (a, b), Zi E Zl+ for i = Zl, ... , Zn is a basis of the space Zl+'
331
1, ... , nand Zi(C) = ei, then •
A basis of the space Zl+ is called a fundamental system of solutions of the equation (1.3) (or of the corresponding system). With the notation of Theorems 1.3 and 1.4 the general solution of the equation (1.3) may be written as A1Z1 + ... + AnZn, the general solution of the equation (1.2) may be written as Zo + A1 Zl + ... + AnZn, where A1,' .. , An are arbitrary (respectively, real or complex) numbers. Remark 1.1. Now let Zi = {Zi1"",Zin}' i = 1, ... ,n, be an arbitrary set of n solutions ofthe equation (1.3). Put: Zll
~(t) =
(t)
Z12(t) Zl n (t)
(the function ~ is called a Wronskian of the system of functions Zl,' .. , zn). Calculate the derivative of the function ~. We have:
0 OXij
Xll
X1j-1
X1j
X1j+1
X1n
Xi-ll
Xi1
Xi-1j-1 Xij-1
Xi-1j Xij
Xi-1j+! Xij+1
Xi-1n Xin
Xi+ll
Xi+1j-1
Xi+1j
Xi+1j+1
Xi+1n
Xn1
Xnj-1
Xnj
Xn j+1
Xnn
Xll
X1j-1
X1j
X1j+1
X1n
Xi-ll
Xi-1j-1
Xi-1j
Xi-1j+!
Xi-1n
0
0
1
0
0
Xi+ll
Xi+1j-1
Xi+!j
Xi+1j+1
Xi+1n
Xn1
Xnj-1
Xnj
Xnj+1
Xnn
By Lemma 10.2.1 ~/(t) = ~l(t) + ... +~n(t), where ~i is the determinant, which is obtained from the determinant ~ by changing its i-th row to the row (Z~i(t), ... , Z~i(t)). Since the functions Zl,"" Zn solve (1.3), the i-th row of the determinant ~i may be represented as a linear combination of the rows of the determinant ~ with the coefficients ail, . .. , ain, respectively. Therefore ~i = aii~ and ~/(t) = tr A . ~(t) (for almost all t E (a, b)), where tr A = all + ... + ann- The 'last relation is a differential equation.
332
CHAPTER 11
The formula (3.1) of Example 8.3.3 gives its solution
where to denotes an arbitrary point of the interval (a, b) (Liouville '8 formula). When we solve a linear equation, we often use the change of variables
y(t) = C(t)x(t),
(1.4)
where (in addition to our notation)
Cln(t))
C2n (t)
,
cnn(t) Assume that the functions Cij, i, j = 1, ... , n, are generalized absolutely continuous and for every t E (a, b) the determinant of the matrix C(t) does not vanish. The equality (1.4) is equipotent to the system
j
Yl(t) = CU(t)Xl(t)
+ Cl2(t)X2(t) + ... + Cln(t)Xn(t),
~2.(.t~.~. ~~~ ~t.).~l.~t~.~. ~~2.(.t~~.2.(~~ .~ ....... ~ ~~~~~).~~(t),
Yn(t) =
Cnl (t)Xl (t) + Cn2(t)X2(t) + ... + cnn(t)xn(t)·
We can apply remarks of Example 10.2.3 and Lemma 10.2.4 to this change. In particular,
yl = Clx +Cxl
(1.5) by the formula (10.2.4). Here
c~n(t)) c~n (t) c~n(t)
Solve (1.5) with respect to Xl and substitute (1.2) and (1.4). We obtain the equation (1.6)
Xl = C-l(AC - CI)x
+ C-1a
(we have repeated in the particular case a calculation which gives in the general case the equation (10.2.8).
Some methods of investigation of equations.
333
Example 1.1. Consider the equation (1.3) but assume in addition that the matrix A is constant (Le., functions Cl'.ij are constant). As is well known (see, for instance, [ER]), there exists a non-degenerate (constant) matrix C such that the matrix J = C- I AC has the Jordan form
o J=
o where
o
0
The change (1.4) with the constant matrix C mentioned transforms the equation (1.3) following formula (1.6) into the equation (1. 7)
X'
= Jx
(here C ' is the zero matrix). The system which corresponds to the equation (1. 7) falls into subsystems of the form:
(1.8)
{
(which correspond to 'boxes' of Jordan form). Its solutions may be found quite simply. The first equation of the system (1.8) contains only one variable Xik and the equation may be solved by the formula (8.3.1). When the function Xik is found we make a corresponding substitution in the second equation of the system (1.8), which now turns out to be an equation with the unique unknown variable Xik+1 of the type studied in Example 8.3.3. Next we pass to the following equation, etc .. This allows us to solve equation (1. 7) and so the initial equation (1.3) too. It is appropriate to notice that for a large number of variables finding of a matrix of Jordan form is often a rather complicated problem. On the other hand in many cases the way outlined turns out to be convenient. For instance, the arguments of Example 10.2.6 follow this path. Finding of a Jordan form, as a rule, is related to the use of complex numbers. To find
CHAPTER 11
334
real solutions of a real equation (1.3) it is often convenient first to find all its complex solutions and then to select the real solutions from amongst them. Example 1.2. With the notation of the (1.6) let the functions Zl = {Cll, ... , Cnl},
. .. , Zn = {Cl n , . .. , Cnn }
constitute a fundamental system of solutions of the equation y' = By of the type (1.3). Then C ' = BC. After the corresponding substitution the equation (1.6) takes the form (1.9)
X'
= C-I(A - B)Cx
+ C-Ia.
In the particular case when A = B we obtain an easily solvable equation x' = C-Ia. So when we find an arbitrary fundamental system of solutions of the equation (1.3) we obtain an easy possibility of solving equation (1.2) with an arbitrary function a. This way of solving non-homogeneous linear equations is called the method of variation of constants. Recall that by Theorem 1.3 to find a general solution of equation (1.2) we need only point out one of its particular solutions Zo and find a general solution of the equation (1.3). 2. Linear equation of order n Consider the equation
(2.1)
y(n)
+ an_l(t)y(n-l) + ... + al(t)y' + ao(t)y + a(t)
= 0,
where real or complex functions ai, i = 0,1, ... , n -1, and a are defined on an interval ( a, b) ( - 00 ~ a < b ~ (0) of the real line, the functions ai are (locally) Lebesgue integrable and the function a is Denjoy integrable. In the case when a == 0 equation (2.1) is called (linear) homogeneous (of order n), and in the opposite case it is called non-homogeneous. The substitution YI = Y and Yk = y(k-l) for k = 2, ... , n transforms the equation (2.1) into the system y~ =
y; =
Y2, Y3,
(2.2) Y~-l = Yn, Y~ = -aO(t)YI - al(t)Y2 - ... - an-l (t)Yn - a(t).
Let us apply the remarks of the previous section. Let Z = {Zl' ... ' zn} be a solution of the system (2.2). It is natural to introduce the definition of a
Some methods of investigation of equations.
335
solution of the equation (2.1), for which the function Zl is considered as a solution. According to results of §1 every solution of the equation (2.1) may be extended on the interval (a, b) and set of all such extensions for a == 0 constitutes a vector space. Theorem 1.4 implies that the dimension of this space is equal to n. Its (arbitrary) basis is called a fundamental system of solutions of the corresponding homogeneous equation and a general solution of the equation (2.1) may be written as Zo + A1Z1 + ... + Anzn, where Zo denotes an arbitrary (particular) solution of the equation (2.1), Zl,"" Zn is a fundamental system of solutions of the corresponding homogeneous equation (i.e., of the equation obtained from (2.1) by rejection of the term a(t)), AI"'" An are arbitrary numbers. Our final remarks in this section concern equations with constant coefficients. Let ao(t) == ao,···, a n-1(t) == an-I, where ao, ... , a n-1 are numbers. The polynomial
is called the characteristic polynomial of the equation
(2.3)
y(n)
+ a n_1y(n-1) + ... + a1Y' + aoy + a(t)
= O.
Let AI,"" As be roots of our characteristic polynomial and k 1, . .. , ks be their multiplicities (k1 + ... + ks = n). The direct verification shows that the functions e >'lt , te>'lt , ...
e A• t te A• t , tk,-1eA,t "e A2t ... , ' , ... , t k . - 1 e A• t
are linearly independent and solve the equation (2.4) This implies that the functions mentioned constitute a fundamental system of solutions of the equation (2.4). We may apply the method of the variation of constants to solve the equation (2.3) (see the end of the previous section). When we seek real solutions of the equation we can first find all its complex solutions and then select the real ones amongst them.
3. Modification of the right hand side. Equations with discontinUous right hand sides The general idea of using of a change of variables for finding of solutions and for investigation of properties of equations (inclusions) consists in a change of the investigation of equation under consideration to the investigation of another equation (inclusion). The same idea of changing an equation to
336
CHAPTER 11
another one is used in the following approach to establishing properties of solutions of an equation with discontinuities in the right hand side, see also [Faf, end of §7]. We have obtained already some results about properties of solution spaces under very general assumptions. In order to use results for investigating a particular equation we need to have the possibility of checking the fulfilment of the axioms of the theory for the solution space of the equation in question. We can do it with the help of Theorems 6.5.1, 8.2.1, 8.4.1 and remarks of Examples 6.5.1,6.5.2,8.3.1,8.3.2 and etc., but in all these results we have as a restriction the requirement of the continuity (respectively, of the upper semicontinuity) of the right hand side f(t, y) in the argument y. Here we study a possibility of the expansion of these results on equations (inclusions) with discontinuities in the space variable. We change an equation (inclusion) with discontinuities to an inclusion with the corresponding continuity and with the same solution space. The proof of the coincidence of two solution spaces may be based on the following simple assertion. Let as before U be an open subset of product ]R x ]Rn. Theorem 3.1. Let F, G : U -+ ]Rn be multi-valued mappings and ttD(G)( {x: x E U, F(x) ~ G(x)}) = O. Then D(G) S;;; D(F). Proof. Let the domain 7r(z) of a function Z E D(G) contain more than one point. By our hypothesis the measure of the set Mo = {t : t E 7r(z), F(t, z(t)) ~ G(t, z(t))} is equal to zero. Since Z E D(G) for every point t of a subset MI of full measure of the segment 7r( z) the (approximate) derivative z'(t) exists and belongs to G(t, z(t)). For every point t E MI \ Mo we have z'(t) E G(t, z(t)) S;;; F(t, z(t)). So z E D(F). In view of the arbitrariness of the function z E D( G) this implies the assertion in question. The theorem is proved. • Corollary. Let the hypotheses of Theorem 3.1 hold and F(x) S;;; G(x) for every point x E U. Then D(F) = D(G). Proof. The additional assumption implies immediately the inclusion D(F) S;;; D(G) (or rather, for its justification we may refer to Theorem 3.1 too). Together with Theorem 3.1 it gives what was required. • Now let F : U -+ ]Rn be an arbitrary multi-valued mapping. For (t, y) E U put F(t, y) = n{ cc(Ft(V)): V is a neighborhood of the point yin Ud. Evidently F(t, y) S;;; F(t, y) for every point (t, y) of the set U. Since for every neighborhood V of the point y we have
then
Some methods of investigation of equations.
n{ cc(Ft(V)): V is a neighborhood of the point y in Ut } <;;; n{ cc( F t (V)): V is a neighborhood of the point y in =
337
Ud
Ft(Y)·
Hence F E Q*(U), see §6.4. By the remarks of §6.4: Lemma 3.1. Let a mapping F : U --t JRn satisfy condition a) of §8.2. Then F E Qd(
({x : x E U, F(x)
t= F(x)})
= O.
Or rather, when we solve any real (physical and etc.) problem we may use this construction in the cases when the last condition is not fulfilled. But then the question remains open, of which relations are between solutions of the modified inclusion and solutions of the initial one. Concerning the estimate itself, Lemma 7.4.1 suggests a convenient method of obtaining it. Example 3.1. Consider the scalar equation y' = f(y), where
f(y) =
{~y
for y for y
~
>
1, 1.
The above construction leads to the consideration of the differential inclusion y' E j(y) with the right hand side
{y} { j(y) = [1,2] {2y}
for y < 1, for y = 1, for y > 1.
By Theorem 1.1 and remarks of §10.2 D(j) E Rcekn(JR x JRn). Denote by Uo the strip JR x (t, 2). By the Mean Value Theorem 4.10.1 D(j, Uo) <;;; Ml (h, 1) (see §7.4) for h(t, y) = y. All functions from Ml (h, 1) are increasing. Therefore the graph of every function Z E D(j) meets the set M = {( t, y): (t, y) E JR x JR, y = I}
= {(t, y): (t, y) = h-1(1)
E
JR x JR, j(t, y) ~ {j(t, y)}}
CHAPTER 11
338
not more than at one point. By Theorem 7.4.4 ILDcJ)(M) = O. Theorem 3.1 implies the coincidence of the spaces D(j) and D(f). Thus
D(f) E Rcekn(IR x JR).
Figure 11.1
Figure 11.2
Our equation is easy to solve explicitly. The formula of Example 8.3.3 represents apart solutions of our equation (and inclusion) in the regions U1 = {(t,y): (t,y) E JR x JR, y
<
I}
and
U2 = {(t,y): (t,y) E JR x JR, y> I}. We can write a general solution of the equation y(t) = ee t for e ~ 0 and
y(t) = {
eet 2 2t
e e
for t ~ -Inc, for t ;?: - In e
for e> 0 (Figure 1l.1). Example 3.2. Consider the scalar equation y' = f(t, y), where
f(t,y) =
{~y
for y ~ t/2 for y > t/2.
The construction mentioned leads to the consideration of the differential inclusion y' E j(t, y) with the right hand side
{y}
j(t, y) =
1
[2y, y] [y,2y] {2y}
for for for for
y ~ t/2, y = t/2 ~ 0, y = t/2 > 0, y > t/2.
Some methods of investigation of equations.
339
As in the previous example, D(f) E Rcekn(lR x lR) by Theorem 8.2.1 and remarks of §8.3. In our case the solution space of the inclusion y' E j( t, y) does not coincide with the solution space of the equation y' = f (t, y): the function zo(t) = ~,where k ~ t ~ 1, solves the inclusion y' E j(t,y), but it does not solve the equation y' = f(t, y), as well as its restriction to every non-degenerate segment lying in 1T(Zo), The complement in D(f) to the set consisting of the function Zo and of its restrictions on arbitrary non-degenerate segments lying in 1T(ZO) and their extensions, coincides with the solution space of the equation y' = f (t, y) (outside the segment Gr( zo) we estimate the behavior of solutions As in the previous example but with a suitable function h). In points of the segment Gr(zo) the space does not satisfy condition (e) (Figure 11.2). 4. Simplest singularities. Existence of solutions Let U be an open subset of the space lR x lR n , h : U -+ lR be a continuous mapping, Z E Rc(U). for every function z E Z let the function
[h- 1 (( -00, c»] n [h- 1 ((c, (0»] and ZU\F E Re(U \ F). Then Z E Re(U), Proof. We need to prove that condition (e) holds if we take as a point appearing in its statement an arbitrary point (to, Yo) of the set F. Take a number rJ > 0 such that 0/((t o, Yo), 'T}) <;;; U. We can fix points (b i , y;) E O((to, Yo), rJ)
n h- 1 (( -00, c»,
i = 1,2, ... ,
in a manner that the sequence {(b i , Yi): i = 1,2, ... } converges to the point (to, Yo). By virtue of the condition ZU\F E Rce(U \ F) and the choice of the points (b i , y;), i = 1,2, ... , there exists a function Zi E Z- such that bi = sup 1T(Zi) and zi(b i ) = Yi' Lemma 6.6.1 implies the existence of a point ai E 1T(Z;) such that
and
(ai, Zi(ai» E 80«to, Yo), rJ)· The set 0 1 « to, Yo), 'T}) <;;; U is compact and contains the graphs of all elements of the sequence a = {zii[a;,b;J : i = 1,2, ... }. By virtue of condition (c) the sequence a contains a subsequence converging to a function z* E Z. By Theorem 3.5.4 (inf 1T(Z*), z* (inf 1T(Z*») E 80« to, Yo), 'T}),
340
CHAPTER 11
to = sup 7r{Z*) and
z*(to) = Yo· Likewise there exists a function z** E Z with
to
= inf7r{z**) z**{to) = Yo,
(sup 7r{z**), z**{SUP7r{z**») E aO{{to, Yo), "'). The point to lies in the interior of the domain 7r{ z*) U 7r{ z**) of the function
z*{t) z{t) = { z**{t)
ift E 7r{z*), ift E 7r{z**),
which belongs to Z. This gives what was required. The theorem is proved. • Theorem 4.1 may be generalized without any essential additional efforts to the following assertion. Theorem 4.2. Let M be an at most countable subset of the real line.
Let F
~
h-1{M) be a closed subset of the set U and ZU\F E Re{U\F). Let
• for every c EM. Then Z E Re{U). Proof. Denote by "( the set of all open subset V of the set U satisfying the condition Zv E Re{V). The set F* = U \ U"( lies in F. The set M* = h{F*) lies in M. Hence it is at most countable. Assume that the set F* is nonempty. Take an arbitrary x E F*. For some c> 0 we have O/{x,c) ~ U. The set Ml = h{O/{x,c)nF*) is compact (see Theorems 2.4.1 and 1.7.8). It lies in M*. Therefore it is at most countable. Its closed subset M2 = [h{O{x, c) n F*)] is at most countable too. By Assertion 2.4.1 the set M2 has an isolated point t. The set {t} is open in M 2. Therefore it meets the set h{O{x, c) nF*), i.e., t E h{O(x, c) nF*) is an isolated point of the set h{ O{x, c) n F*). Theorem 4.1 and the definition of F* imply that such a point cannot exist. Our assumption is false, F* = 0, . U
= U"(.
Referring to Lemma 6.5.2 completes the proof. • It is helpful to compare the proved assertion with results of Chapter 7, where between other we have pointed possibilities of the verification of the fulfilment of condition (c). Consider a simple example. Example 4.1. Let Z E Rce(U). Let F be an at most countable closed subset of the set U and ZU\F E Rce{U \ F). Show that Z E Rce{U). By Corollary of Theorem 7.6.7 and by Theorem 7.6.6 (with Cs(U) as '11) Z E Rc(U). By Theorem 4.2 (with h(t, y) == t) Z E Re(U).
Some methods of investigation of equations.
341
When we consider an equation y' = f (t, y) with singularities in the right hand side, sometimes it is convenient to change the right hand side (see, for instance, §3). If in a point of singularity the right hand side is not defined then we may often define' it rather arbitrarily. We mean that results of further considerations will not depend on the way in which we define the right hand side, see Example 8.2.1 and Example 4.2 below. Finally, as rule, we can change the consideration of the equation y' = f (t, y) to the consideration of the inclusion y' E j(t,y), where j(t,y) = U(t,y)}, if the value f(t,y) is defined, and j(t,y) = 0 in the opposite case. In many cases this makes us free from a necessity of any additional discussion of the presence of such singularities of the right hand side of the equation. Example 4.2. Consider the scalar equation
ty
y'
Outside the closed set
its right hand side is continuous. Therefore on the set ]R2 \ F the solution space of the equation satisfies conditions (c) and (e). By Theorem 7.6.6 and remarks of Example 4.1 the solution space satisfies conditions (c) and (e) on the entire plane ]R2. Example 4.3. Let Zl E Rp(U), u E ]Rn, Ilull =I=- 0, h(t,y) = (u,y), E > 0, Zl ~ Ml (h, E) (see §7.4). Let F be a closed subset of U, (Zl)U\F E Rce(U \ F). Let for every t E ]R the set Ft be at most countable. Show that Zl E Rce(U). By Lemma 7.4.2 the set W = M1(h,E) is closed in Cs(U). By Corollary of Theorem 7.6.7 and by Theorem 7.6.6 the space Zl satisfies condition (c). Referring to Theorem 4.2 completes the consideration. Example 4.4. Let V be an open subset of the space ]Rn. Let a function f : V ~ ]Rn be continuous outside an at most countable closed subset F of V. Let u E ]Rn, Ilull =I=- 0, E > and (f(y),u) > E for every point y E V. In this case the solution space of the equation y' = f (y) satisfies all restrictions of the previous example. Example 4.5. By remarks of Example 4.4 the solution space of the two-dimensional system
°
X' {
y
,
= =
cosy,
xy
.
sln2
1 -
X
. 2 1 + sln
Y
CHAPTER 11
342
satisfies locally conditions (c) and (e). By Corollary of Theorem 7.4.4 and by Lemma 6.5.2 it satisfies conditions (c) and (e).
5. Collage of spaces Let U be an open subset of the space lR x lR n , h : U - t lR be a continuous mapping. For Z E Gs(U) consider the function
in U to the space Z* *a Z**. Proof. Let K E expc U. Let B be an infinite subset of Nand Zi E Z; *a, Zt* for i E B. Our aim is to show that there exists a subsequence of the sequence {Zi: i E B} converging to an element of the space Z* *a Z**. The following three cases are possible. 1. Let the set Bl = {i: i E B,
Some methods of investigation of equations. By Theorem 3.5.1 supn(z*) 3.5.4
= limi-->oo,iEA3 O:"i = inf n(z**).
343 By Theorem
By Theorem 1.7.6 the function
z*(t) z(t) = { z**(t)
for t E n(z*), for t E n(z**),
1f(z) = n(z*) U 1f(z**), is continuous. By Theorem 3.5.5 the sequence i E B 3 } converges to z. By Theorem 3.5.4 h(supn(z*),z*(supn(z*)))
= t--> .lim ai
=
a
{Zi :
= h(infn(z**),z**(infn(z**))).
00
By virtue of the condition Z*, Z** E M (h) the function z belongs to the space Z* *a Z**. II. Let the set Bl = {i: i E B, ipzi(supn(zi)) :::; ad be infinite. Then Zi E (ZnK for every i E B 1 . We obtain what was required from the convergence of the sequence {Zt: i = 1,2, ... } in U to the space Z*, from Theorem 3.5.4 and from the continuity of the function h. III. The case when the set Bl = {i: i E B, ipzi (inf n(zi)) ;? ad is infinite, is similar to II. The theorem is proved. • Let us return to the definition. Remark 5.1. If Zl, Z2 E R(U) and r E JR, then Zl *r Z2 E R(U). If, in addition, the spaces Zl and Z2 satisfy one of conditions (c), (e), (u), (k), (p), (q) or (n) then the space Zl *r Z2 satisfies the corresponding condition too. For conditions (e), (u), (k), (p) and (q) this is obvious. For condition (c) this follows easily from Theorem 5.1 (when we apply it to the stationary sequences ai == r, Zt == Zl and Zt* == Z2)' For condition (n) this follows from Lemma 8.3.1. Remark 5.2. Let F 1 , F2 : U --+ JRn be multi-valued mappings. Let Zl = D(Fl) and Z2 = D(F2)' Then Zl *r Z2 = D(F), where
Fl(X) { F(x) = ~(x)
if h(x) < r, if h(x) > r, if h(x) = r.
The just introduced notion may be used to continue the investigation of simplest singularities started in the previous section. We have:
CHAPTER 11
344
Lemma 5.1. Let
d = lim ai = lim bi , ZI E Rcp(U), Z2 E Rc(U), ZI, Z2 E M(h) and 2---+00 1.---+00 Z;* = (ZI *ai Z2)*bi ZI for i = 1,2, .... Then the sequence {Z;* : i = 1,2, ... } converges in U to the space ZI as i ---t 00. Proof. The proof reduces to referring to Theorem 5.1, because in this case (ZI *d Z2) *d ZI = ZI' • Theorem 5.2. Let ZI E Rcp(U), Z2 E Rce(U) (respectively, Z2 E RcedU)), ZI, Z2 ~ M(h), F be a closed subset of the set U, the set h(F) be at most countable, and (ZI)U\F E Re(U \ F) (respectively, (Zdu\F E Rek(U \ F)). Then ZI E Re(U) (respectively, ZI E Rek(U)), Proof. 1. If the set F is empty, our assertion is obvious. II. Assume that the set h(F) consists of one point d only. Choose points al < a2 < ... < d < ... < b2 < bi such that d = limi->oo ai = limi->oo bi and construct the sequence a = {Z;*: i = 1,2, ... } according to Lemma 5.1. By Lemma 5.1 the sequence a converges in U to the space ZI. Now we obtain what was required from Theorems 7.2.1 and 8.1.1. III. We can deduce the general case from II as in the proof of Theorem 4.2 (with the addition ofreferring to Lemma 8.1.3). This remark completes the proof. • Corollary. Let ZI E Rp(U), F be a closed at most countable subset of the set U and (ZdU\F E Rcek(U \ F). Then ZI E RcedU). In order to obtain the corollary notice that the condition ZI E Rc(U) follows from the Corollary of Theorem 8.6.7 and from Theorem 7.6.6 (as in Example 6.1). Next, consider the function h(t, y) == t and take as Z2 the solution space of the equation y' = O. • Example 5.1. By the Corollary of Theorem 5.2 the solution space of the equation of Example 4.2 satisfies the Kneser condition. We leave to the reader the possibility of analysing independently, but with help of Theorem 5.2, the situations described in the other examples of the previous section. 6. Simplest singularities. Approximations We shall keep the notations of the previous section. In Theorem 5.2 and in its Corollary we did not exhaust all possibilities of the situation described. The ideas used may help in the proof of the presence of other properties of solution spaces (besides (e) and (k)). One of main steps of the arguments in the previous section was the passage to the limit in the space Rc(U), It may be helpful in many other cases too.
Some methods of investigation of equations.
345
A. Let Zl E Rcp(U), Z2 E Rc(U), Zl, Z2 E M(h). Let X be a closed subset of the set U, the set h(X) be (at most) countable, K be a compact subset of set U and E > O. The set h(X n K) is compact and at most countable: h(X n K) = {Xi: i = 1,2, ... }, where the points X; defined may not be for all indices i, but Xi ¥ Xj if i ¥ j. For i = 0,1,2, ... construct inductively spaces Z;* and numbers 8i > O. For i = 0 put Z; = Zl, 80 = E. Let the spaces Z; and 8j be constructed for j = 0, ... ,i. Construct Z;*+l and 8i+l' For 8 > 0 consider the space Z*(8) = (Z;* *Xi- 8 Z2) *xiH z;'- By Lemma 5.1 the sequence {Z*(2- j ) : j = 1,2, ... } converges in U to the space Z;*. Therefore with the notation of §7.7 for some 8i+ 1 > 0 we have a((Zt)K, (Z*(8i+d)K) < 2-;-l E. Put Z;'+l = Z*(8i+1)' The family {O(x;, 8;): i = 1,2, ... } of open subsets of the real line covers the compactum h(X n K). Pass to a finite subcover {O(x;, 6;) : i=l, ... ,p}.
Since (see §3.4)
(6.1)
a((ZdK' (Z;)K) ~ a((Z;)K' (Z;)K)
+ ... + a((Z;_l)K, (Z;)K)
<2- 1 c+···+2- P E<E,
the space Z = Z; approximates quite well the space Zl (see §7.7). By virtue of the representation
we can apply the remarks of the previous section to establishing the properties of the space Z. In addition, X n K ~ h- 1 (u{O(x;, 6;): i = 1, ... ,p}). B. Let Zl E Rcp(U), Z2 E Rc(U). For i = 1, ... , io let functions hi : U --7 IR. be continuous, Zl, Z2 E M(h i ), Xi be closed subsets of the set U. Let the set h;(X;) be at most countable, X* = U{X; : i = 1, ... ,i o}, K E expc U and E > O. Construct sequentially spaces Z;**, i = 0,1, ... ,io. Put Z;* = Zl' Let the space Z;**, i < i o , be constructed. Construct the space Z;*;l with Z;** (as Zd, hi (as h), Xi (as X) and 2- i - 1 c (as E) according to A. Analogously to (6.1) we obtain the estimate a((ZdK' (Zi:*)K) < c. Hence the space Z;*o* approximates well the space Zl' Our definition of the space Zi:* gives the possibility of applying the remarks of the previous section to the establishment of its properties.
346
CHAPTER 11
c. The construction of B uses an induction with the common step described in A. We can continue it to a transfinite induction (see §§1.8 and 9.4). The scheme of arguments A-C may be used in various situations. Point one of them. Theorem 6.1. Let F : U -+ ~n be a multi-valued mapping, D(F) E Rc(U). Let a multi-valued mapping Fo : U -+ ~n satisfy locally the Davy conditions, P E expc(U). Let, be a family of open subsets of U. Let the mapping Flv satisfy for every V E ,. Let E = U \ (U,). Let the set E may be represented as the union of closed subsets E i , i = 1, ... , i o . Let single valued functions hi : U -+ ~n, i = 1, ... ,io, be continuous, D(F), D(Fo) E n{M(h i ): i = 1, ... ,io}. Let the set hi(Ei ) be at most countable for every i = 1, ... , io. Then there exist a neighborhood OP ~ U of the compactum P and a sequence {Zj: j = 1,2, ... } ~ Rceu(OP) converging in OP to the space D(F, OP). Proof. The proof follows the induction on the number i) = 0,1, ... ,io satisfying the condition: if Ei n P -I 0 then i ~ i) for all arbitrary U and P E expc U simultaneously. 1. If i l = then En P = 0. By Lemma 2.4.2 O,(P, 8) ~ U \ E for some 8 E (0, (0); moreover, the set PI = O,(P, 8) is compact. Put OP = O(P, 8). Our assertion follows from the constructions of §8.2 (we leave to the reader as an easy exercise the construction of a majorant of the right hand side: use the compactness of the set [OP] ~ U \ E). II. Let the assertion is proved for i l = j. Prove it for i) = j + 1. By the definition of i) we have (U{Ei : i = j+2, ... ,io})np = 0. By Lemma 2.4.2 O,(P, 8) ~ U \ (U{Ei: i = j + 2, ... , io}) for some 8 E (0, (0); moreover, the set O,(P, 8) is compact. Put OP = O(P, 8). Take an arbitrary "1 = 1,2, .... Repeat the construction of A with ZI = D(F), Z2 = D(Fo), h = hj+I' X = Ej+I' K = [OP] and c = 2-1). Let with the notation of A ZI) = Z and GI) = U{ O(Xi' 8i ) : i = 1, ... ,p}. Let W ~ OP be an arbitrary neighborhood of the set Ej+1 n OP with the closure lying in hj~1 (GI)). Since the mapping Fo satisfies locally the Davy conditions, D(Fo) E [Rceu(U)]. By the inductive hypothesis, by Lemma 7.7.2 and by Theorem 7.7.1 D(F,OP\[W]) E [Rceu(OP\[W])]. By remarks of the previous section, A, Lemma 7.7.2 and Theorems 7.6.4 and 7.7.1 this implies that (ZI))oP E [Rceu(OP)]. By Theorem 7.7.2 the sequence {(ZI))oP : "1 1,2, ... } converges in OP to the space D(F,OP). Hence D(F,OP) E [Rceu(OP)], which means the validity of the assertion in question.
°
Some methods of investigation of equations.
347
III. When i l = io in II we obtain what was required. The theorem is proved. • Consider a near but a slightly different situation. D. Let F : U -+ IR,n be a multi-valued mapping. Let a function h : U -+ IR, be continuous and have continuous derivatives. Let ~~ (x) > 0 and UI [}[}h (x) + ... + Un[}fJh (x) > 0 for every x = (t,XI,'" ,x n ) E U and Xl Xn U = {UI,"" Un} E F(x). Let Y be a compact subset of the real line Jt(Y) = O. Let Fo(x) = (~~(X))-l{:xh, (x), ... , (x)) for x E U. By Lemma 10.2.1 the function cpz(t) = h(t, z(t)) is generalized absolutely continuous for every function z E D(F). At almost every point its approximate derivative is positive. By Lemma 4.9.1 the function cpz is nondecreasing. On every (non-degenarate) segment the function
:x:
ah
cp~(t) = at (t, z(t))
+ (ah at (t, z(t)) )-l((ah aXI (t, z(t)) )2 + ...
+ (::n (t,z(t)))
2) > 0,
we have D(Fo) E M(h). By Theorem 2.6.7 for every b > 0 the set O(Y, b) may be represented as the union of a set 1 of its connected components. Since here 1 is an open cover of the compactum Y consisting of pairwise disjoint sets meeting Y, the set 1 is finite: 1 = {(ai, bi ): i = 1, ... ,p}. Put Zl = D(F), Z2 = D(Fo) and Z(b) = ( ... ((Zl *al Z2) *b 1 Zd *a2 ... *a p Z2) *b p Zl' Lemma 6.1. Under the assumptions of D let the mapping F satisfy locally the Davy conditions. Then the generalized sequence {Z (b): b > O} converges in Rc(U) to Zl as b -+ O. Proof. The membership Z(b) E Rc(U) follows from results of §§8.2, 8.3 and §7. Now take an arbitrary sequence {b j : j = 1,2, ... } of positive number converging to zero, a compactum K ~ U and functions Zj E (Z(bj))K' j = 1,2, .... Let
CHAPTER 11
348
point only, then the membership z E Zl is obvious. Consider the opposite case. By Lemma 10.2.1 (ZdK' (Z2)K, (Z(8))K ~ Ml (h, E), where E = inf~~(K) > O. Since by Lemma 7.4.2 the set M1(h,E) is closed in Cs(U), z E Ml(h,E). We have seen in §6.3 that the space Z consists of absolutely continuous functions. Let J = {t: t E 7I"(z), (t, z(t)) ~ h-l(y)}. By Lemma 7.4.3 the set J is an (open) subset of full measure of the segment 71"( z) and for almost all t E 7I"(z) we have z'(t) E F(t, z(t)). By Lemma 1.7.1 the lemma is proved. • Lemma 6.2. Let under the assumptions of D , be a family of open sub-
E,
sets ofU. For every V let the mapping Flv satisfy the Davy conditions. Assume that for every point of the set E = U \ it has a neighborhood W ~ U, a continuous function g : W --+ lR with continuous derivatives, and a number c ~ 0 such that og og og -(x) + Ul-(X) + ... + un-(x) ~ c at OXI OXn for every x E Wand {Ul, ... ,un} E F(x) U {Fo(x)}. Let Jt(g(E n W)) = 0 and Jt(h(E)) = O. For every t E lR let the set E t contain no non-trivial connected subset. Let a set Uo ~ U be open, its closure be compact and lie in U. Let spaces (Z(8))uo' 8 > 0, be constructed according to D and Y = h([UoJ nE). Then the generalized sequence {(Z(8))uo : 0 > O} converges in the space Rc(Uo) to D(F, Uo) as 0 --+ o. Proof. I. By the Corollary of Theorem 7.6.7 D(F) E Rs(U). By Theorem 8.2.1 D(F, V) E Rc(V) for every V E ,. By Theorem 7.6.5 and by the results of §7.4 D(F, W) E Rc(W) for every set W satisfying the hypotheses of the statement. By the Corollary of Theorem 7.6.4 D(F) E Rc(U). Now the membership (Z(o))uo E Rc(Uo) follows from remarks of the
U,
previous section. II. By Lemma 6.1 the generalized sequence {(Z(o))vnuo: 0 > O} converges in the space Rc(V n Uo) to D(F, V n Uo) as 0 --+ 0 for every V E ,. By Corollary 1 of Theorem 7.3.1 {(Z(8 j ))uo : j = 1,2, ... } E s(Uo) for every sequence {OJ: j = 1,2, ... } of positive numbers converging to zero. By results of §§7.4 and 7.5 the sequence {(Z(8 j ))w: j = 1,2, ... } converges in Rc(W) to D(F, W) for every W ~ Uo satisfying the hypotheses of the statement. Referring to Theorem 7.6.4, Lemma 1. 7.1 and results of • §7.7 completes the proof. Theorem 6.2. Under the hypotheses of Lemma 6.2 D(F, Uo) E [Rceu(Uo)]· Proof. The proof reduces to referring to Lemma 6.2 and remarks of §7.7. • Notice one more possibility of construction of 'good' approximations of solution spaces.
Some methods of investigation of equations.
349
Let H ~ ~n. Denote by M3(H) the set of all functions z E Cs(U) satisfying the condition z(tt:;(s) E H for any two different points s, t E 7f(z). Lemma 6.3. Let a set H ~ ~n is closed. Then the set M3(H) is closed in the space Cs (U). Proof. The proof is analogous to the proof of Lemma VII.4.2. • Theorem 6.3. Let F : U --t ~n be a multi-valued mapping, , be a family of open subsets of the set U. Let for every V E , the mapping Flv satisfy the Davy conditions. Assume that for every point of the set E = U \ (U,) there exists its neighborhood W ~ U such that
(6.2) where H = cc(F(W)), and for every t E ~ the set E t does not contain non-trivial connected subset. Then D(F) E [Rceu(U)] n Rc(U), Proof. I. By Theorem 4.10.1 and 7.6.5, Lemma 6.3, and Corollary of Theorem 7.6.7 D(F) E Rc(U), II. Let Q ~ U and F( Q) i= 0. Associate to Q an arbitrary vector u(Q) E F(Q). For k = 1,2, ... and x = (io,i1, ... ,i n ) E Nn+1 denote by Qk", the (n + 1)-dimensional cub
For x E Qk", put
_ GkAx) -
F(x) if Qk", ~ U" U(Qk",) if Qk", Cl U, and the cub Qk", lies . an open set W · f' m , satIs ymg (62) . ,
1
{O} in other cases .
X
Now for k = 1,2,... and x E U put Fk{X) = cc(U{ Gk",(x) E Nn+l, x E Qk"'})' The mapping Fk : U --t ~n satisfies locally the
Davy conditions. The application of constructions of §8.2 to the mapping Fk leads to the membership D{Fd E [Rceu(U)]. III. Show that the sequence {D(Fd: k = 1,2, ... } converges in U to the space D(F). By Theorem 7.6.4 it is sufficient to prove the convergence locally. IV. If x E U" then O,(x,c:) ~ U, for some c: E (0,00). Therefore Fklo(X,E) = Flo(X,E» beginning with some k = k o. By I and Remark 7.6.1 the sequence {D(Fk , O(x, c:)): k = 1,2, ... } converges in O(x, c:) to D(F, O(x, c:)).
350
CHAPTER 11
V. If x E E, then there are an open set W ~ U satisfying (6.2) and f E (0, (0) such that O(x, 2f) ~ W. By Theorem 4.10.1 D(F, O(x, f)) ~ M 3 (H), where H = cc(F(O(x, 2f))). We have D(Fk' O(X,f)) ~ M 3 (H), beginning with some k = ko. Now the convergence of the sequence {D(Fk,O(X,f)): k = 1,2, ... } in O(x, f) to the space D(F, O(x, f)) follows from Theorems 7.3.1 and 7.5.4 and Lemma 6.3. VI. By II-V and remarks of §7.7 the theorem is proved. • Example 6.1. Let F : U -+ ffi.n be a multi-valued mapping, , be a family open subsets of U. Let for every V E , the mapping Flv satisfy the Davy conditions. Let the projection of the set E = U \ (U,) in the first factor of the product ffi. x ffi.n be at most countable and for every t E ffi. the set E t do not contain non-trivial connected subsets. Then by Theorem 6.3 D(F) E [Rceu(U)] n Rc(U). Example 6.2. Let U = ffi.xUo . Let a multi-valued mapping F: U -+ ffi.n do not depend on the first factor (so we can understand it as a mapping from Uo to ffi.n and use the corresponding language, see also §XV.1). Assume that , is a family of open subsets of the set Uo, for every element V of which the mapping Flv is upper semicontinuous, its values are nonempty, compact, and convex. Assume that the set E = Uo \ (U,) is at most countable and for any point of it there exists a neighborhood Wo such that 0 ~ cc(F(Wo)). Show that D(F) E [Rceu(U)] n Rc(U). In order to have the possibility of using Theorem 6.3 we need to point a set W with the properties mentioned in the statement of the theorem. Take ffi. x Wo as W (the description of the set Wo is given above). By Theorem 2.4.2 there exist a linear fuctional f : ffi.n -+ ffi. and numbers p < q such that (with the notation of (6.2)) f(H) ~ (-oo,p] and f(O) ~ [q, (0). Here f(O) = 0 and therefore p < o. Put h(t, y) = - f(y) for (t, y) E W. Since f(H) ~ (-oo,p], we have: M3(H) ~ M 2 (h, -p). The set ffi. x (E n Wo) is at most countable with respect to the space M 2 (h, -p). Hence it is at most countable with respect to the space M 3 (H) too. We now have the possibility of refering to Theorem 6.3. Under our assumptions we have the autonomy of the approximations (see §XV.1). 7.
Optimization
As before let U denote an open subset of the product ffi. x ffi. n . Let a space Z ~ Cs(U) satisfy condition (c), K E K 1 , K 2 E exp K. Let G : Z -+ ffi. be a continuous function.
expc U,
Some methods of investigation of equations.
351
Under our assumptions the set Z K is compact. The set A = { Z : E K 1 , z(SUP7r(z» E K 2 } is closed in ZK. Therefore the set A is compact. Assume the set A is non-empty. By Remark 2.3.4 there are Z1, Z2 E A such that G(zd = inf G(A) and G(Z2) = sup G(A). In applications Z corresponds to a (physical, technical, economic, etc.) process (which may be described, for instance, by a system of ordinary differential equations), G is a 'payment' for the choice of a particular solution of the corresponding problem. The solutions Z1 and Z2 are 'optimal' in a meaning that Z1 is 'cheapest', which may correspond, for instance, to the minimum of expenditures and Z2 is 'most expensive', which may correspond, for instance, to the maximum of profits. Thus the possibility of optimization, i.e., of the choice of a solution, which is optimal in the above meaning, turns out to be directly related with the fulfilment of condition (c). Methods of verifying it were discussed above. Consider the following general situation. Example 7.1. Let the process in question be described by spaces Z1, Z2 E R~(U), Z1 U Z2 ~ Zo E R(U) and a function G : Zo ---t lR be continuous. Assume that we have the possibility of making one switching from Z1 to Z2, i.e., solution z of the problem under consideration may be defined as a function satisfying the conditions zl[a,c] E Z1 and Zl[c,b] E Z2 for some c E 7r(z) = [a, b]. This corresponds to the situation where we consider the space Z = Z1 * Z2 = U{Z1 *r Z2: r E R} with h(t,y) = t in the definition of Z1 *r Z2 (see §5). By Theorems 5.1 and 1.7.8 Z E R~(U). Therefore this optimization problem is solvable.
z
E ZK, z(inf7r(z»
Example 7.2. Let a function g : U
G(z)
=
---t
lR be continuous. Let
J
get, z(t»dt
7r(Z)
for z E Cs(U). The continuity of the function G on Cs(U) follows from Theorems 3.3.1 and 3.7.2. Thus the function G may be taken as a 'payment' in an optimization problem for every space Z ~ Cs(U). Example 7.3. Let U ~ (a, b) x lRn. Let a function cp : (a, b) ---t lRn be locally Lebesgue integrable. The estimate
lJ
(cp(t), z(t»dt
(z)
~ Ilzll
J 1T(Z)
Ilcp(t)lldt
352
CHAPTER 11
(in the first integral we have the scalar product) and Theorem 4.5.1 imply easily the continuity of the function
on Gs(U) (see the proof of Theorem 3.7.2). Example 7.4. Let U ~ (a, b) x IRn. Let a function 'IjJ : (a, b) --t IRn be locally absolutely continuous. Let a space Z ~ Gs (U) consist of generalized absolutely continuous functions and
w(z) = j ('IjJ(t) , z'(t))dt 7r(Z)
for z E Z. By virtue of the representation (3
w(z) = (w(f3),z(f3)) - (w(a),z(a)) - j('IjJ'(t),Z(t))dt, a
where [a, f3] = 1T(Z), by Lemma 10.2.2, remarks of the previous example, Theorem 3.5.4, and remarks of Examples 2.1.3 and 2.1.2 the function \]I is continuous on Z. Every function G of the previous examples satisfies the following additional condition: (7.1)
if z E Z and c E [a, b] = 1T(Z), then G(z) = G(zl[a,c])
+ G(ZI[c,b])'
In connection with these observations notice: Remark 7.1. Let Z E R(U), a function h : U --t IR be continuous. Let with the notation of §8.7 Z ~ M(h), r E IR, Zl = ZIUnh-1((-oo,rJ)' Z2 = ZIUnh-1([r,oo))' Let a mapping G : Z --t IR satisfy (7.1) and for i = 1,2 the mapping Glz; be continuous. Then the mapping G : Z --t IR continuous too. Remark 7.2. Let Z E R(U), 'Y be an open cover of the set U, a mapping G satisfy (7.1) and for every V E 'Y the mapping Glzv be continuous. Then mapping G : Z --t IR is continuous too. 8.
Control
Let U be an open subset of the product IR x IR n , A be a metric space with a countable base and F : U x A --t IR n be a multi-valued mapping. Consider the differential inclusion (8.1)
y' E F(t, y, a),
Some methods of investigation of equations.
353
where a E A play the role of a parameter. A function cp defined almost everywhere on a segment [a, b]' -00 < a < b < 00, of the real line and taking values in A is called a control. The substitution a = cp(t) transforms the inclusion (8.1) into the inclusion (8.2)
y' E F(t, y, cp(t)),
which does not contain parameters. We can consider the right hand side of the inclusion (8.2) defined on the entire set U, by putting it equal to 0 for values of arguments where the previous argument is not applicable. A pair (z, cp), where z solves the inclusion (8.2), is called a solution of the control problem (8.1). In applications, we meet situations, where we consider a process described by an inclusion (8.1) and where we have the possibility of changing values of the parameter a. That is, we can change conditions of the evolution of the process, for instance, when in a technical device we have 'rudders' and their positions control the working of the device. Often we do not consider all the controls, but restrict consideration to a class of controls and select 'admissible' controls. Usually we consider measurable controls. In addition, the current state of a system may impose restrictions on possible values of control, which in our language my be described by a multi-valued mapping ~ : U ---* A. Solution of the control problem for (8.1) with restrictions ~ is a couple (z, cp), where z solves the inclusion (8.2) and a (measurable) control cp satisfies for almost all t E 7f(z) the condition (8.3)
cp(t)
E ~(t,
z(t)).
Respectively, it is sufficient to assume that the function F is defined on the set P = {(t,y,a): t E JR., y E Ut, a E ~(t,y)} only. Here we will not discuss the control problem in exhaustive detail, bat we will watch for the first components of solutions, i.e., for the function z. We mean that we are able to point to a corresponding control for it. By misuse of language such a function z is also called a solution of the control problem (8.1), (8.3). Denote by Z the set of such solutions z. It is not difficult to see that Z E Rpq(U). Hence we can hope to apply our tools to the investigation of the space Z. For (t, y) E U put F*(t, y) = U{F(t, y, a) : a E ~(t, y)}. Evidently Z ~ D(F*). On the other hand, if z E D(F*) then for almost every t E 7f(z) there exists a = cp(t) E e(t, z(t)) such that z'(t) E F(t, y, a). So the function z is a solution of the control problem. But a corresponding control cp need not be admissible. The space D(F*) will consist only of solutions of the control problem (8.1), (8.3) in the precise sense (with admissible controls) When we impose additional conditions on F and
e.
354
CHAPTER 11
Theorem 8.1. Assume that with the above notation: (8.4) for every t E IR the mapping FI{t}xGr~, with nonempty compact values is upper semicontinuous; and if a segment I then
~
(8.5) the mapping
IR and a compactum K
~IIxK
~
IRn are such that I x K
~
U
satisfies conditions (4.13.1-2),
(8.6) for every compactum X ~ K x A and for every open subset V of the space IRn the set {t: tEl, F(({t} x X) np) ~ V} is measurable. Then Z = D(F*). Proof. I. It is convenient to start the argument with the proof of the assertion with the additional assumption that the mapping F is defined on
the entire set U x A and for every t E IR the mapping is upper semicontinuous.
FI{t}xU,XA
with nonempty compact values
The problem reduces to the proof of the existence of a measurable control cp corresponding to an arbitrary function z E D(F*). It is sufficient to prove the existence of such a control locally for a small part of the domain of the function z, i.e., we can assume in addition that Gr(z) ~ I x K ~ U, where I is a segment of the real line and K is a compactum lying in IRn. Let I = [a, b], where -00 < a < b < 00, and E > O. By Theorem 4.13.2 there exists a closed subset Hi of the segment [a, b] such that p,( Hd ~ b - a - ~ and the mapping ~ IH, x K is upper semicontinuous. By Theorem 1.7.8 the set B = ~(Hi X K) is compact. By Lemma 4.13.2 the mapping FI1x(KxB) satisfies condition (13.2) of §4.13. Therefore there exists a closed subset H2 of the segment [a, b] such that P,(H2) ~ b - a - ~ and the mapping FIH2 xKxB is upper semicontinuous. By Luzin's theorem 4.13.7 there exists a closed subset H3 of the segment [a, b] such that P,(H3) ~ b - a - ~ and the function z /lH3 is (defined and) continuous. Let H = Hi n H2 n H 3. By estimates of §4.1 p,(H) ~ b - a-E. Let GH(t) = {,B: ,B E ~(t,z(t)), Zl(t) E F(t,z(t),,B)} for t E H. Show that the mapping G H is upper semicontinuous. We have the possibility of using Theorem 2.5.1 here. Let {ti: i = 1,2, ... } ~ H, ti -+ to E H and ,Bi E GH(ti) for i = 1,2, .... The set B is compact and contains all elements of the sequence "y = {,Bi: i = 1,2, ... }. So we can assume in
Some methods of investigation of equations.
355
addition that the sequence, converges to a point f30 E B. By virtue of the upper semicontinuity of the mapping FIHXKXB, of the continuity of Z and Z'IH and of Theorem 2.5.1 we have:
Z'(t o) = lim z'(t i ) E lim topsUpF(ti' z(t i ),f3i)
~
F(to, z(to),f3o).
i---+oo
'l---+OO
Therefore f30 E GH(to). By Theorem 2.5.1 this means the upper semicontinuity of the mapping G H . In view of the arbitrariness in the choice of c > 0 Theorem 4.13.2 implies the measurablity of the mapping G : I ---+ A defined by the formula G(t) = {f3: f3 E ~(t, z(t)), z'(t) E F(t, z(t), (3)}. By Theorem 4.12.3 there exists a measurable function cp defined almost everywhere on the segment [a, bJ such that cp(t) E G(t) for almost all t E [a, bJ. Evidently cp is a control corresponding to z. Hence z E Z. II. Let us now pass to the general case. Fix two different elements Ul and U2 of the space IR n and for i = 1,2, (t, y) E U and 0: E A, put: F .( t
t, y, 0:
)
= {F(t, y) U {u;} {u;}
if 0: E ~(t, y), if 0:
tt ~ (t, y).
Evidently the mapping Fi satisfies restrictions imposed in I. Let z E D(F*). By I for i = 1,2 there exists a measurable function CPi which is defined almost everywhere on the segment 7r(z) and such that z'(t) E Fi(t, z(t), CPi(t)) for almost all t E 7r(z). The set M = {t: t E 7r(z), z'(t) = ur} is measurable. By Theorem 4.2.3 the function
cp(t) = {CP1(t) CP2 (t)
if t E 7r(z) \ M, if t E M,
is measurable. Evidently z'(t) E F(t, z(t), cp(t)) for almost all t E 7r(z). Hence the function cP is a measurable control corresponding to z. The the• orem is proved. Remark 8.1. In Theorem 8.1 we have imposed on the mappings F and very rigid conditions. Notice that our estimates of sets of singularities are applicable in the situation under consideration too. They immediately expand the domain of application of Theorem 8.l. Theorem 8.1 and Remark 8.1 mean that methods of previous sections are applicable to the discussion of the control problem not only on the level of belonging of the solution space to the set Rq(U). A valuable part of our theory turns out to be applicable to inclusions (equations) with control. We often consider the problem of the finding of an optimal control, i.e., the discussed in previous section optimization problem in the connection
e
356
CHAPTER 11
with the control problem. We have seen in §7 the role in this question the fulfilment of condition (c) play. In the previous chapters we saw how to prove the fulfilment of this condition for equations and inclusions of various types. Example B.l. Let A be a convex compact (nonempty) subset of the space ~m (m = 1,2, ... ), U ~ (a, b) x ~n, a function 10 : U -+ ~n satisfy locally the Caratheodory conditions, elements of the matrix
B(t) = (
b1.1.(.t) bn1 (t)
be locally Lebesgue integrable on the interval (a, b) and
I(t, y, a) = lo(t, y)
+ B(t)a
for (t,y) E U, a E A. In this case the function f* satisfies the Davy conditions. Therefore the solution space of the control problem for the equation y' = I (t, y, a), a E A, belongs to Rcekn(U) (see Theorem 8.1, §§8.2 and 8.3).
CHAPTER 12
EQUATIONS AND INCLUSIONS WITH COMPLICATED DISCONTINUITIES IN THE SPACE VARIABLES
Our aim here is not an exhaustive account of the immense question mentioned in the chapter's title. We show in this chapter which part of our tools may be used for the expansion of the theory to equations and inclusions with complicated discontinuities in right hand sides in space variables. Caratheodory conditions select the time, allow complicated discontinuities in it in right hand sides, and keep only the measurability. In this chapter we show how to include in the general theory equations, in which such discontinuities are also allowed in space variables and in various combinations amongst them, which makes the picture more involved. We use the Scorza-Dragoni theorem and other results of §4.13 in the investigation of right hand sides of equations and inclusions, preparing the application of topological aspects of our theory. A.F.Filippov has proposed a well known approach to the investigation of equations with discontinuities in space variables, see [Faf]. Tallying the development of the corresponding conception and following the ideology of the 'sliding mode', he proposed to change such an equation y' = J (t, y) into a differential inclusion y' E F( t, y) satisfying the Davy conditions. The right hand side of the inclusion broadens the right hand side of the initial equation: J(t, y) E F(t, y). (We are not very precise here. See [Faf] for precise definitions.) He proposed considering every solution of the inclusion y' E F(t, y) as a solution of the initial equation. But first, such a broadening need not exist. Secondly, if we follow a natural (our) definition of a solution, such a passage may lead to a modification of the solution set, although this does not take place every time, see Theorem 11.3.1 and its Corollary. Simples examples of a modification of the solution set under such a passage may be constructed at the level of a scalar equation y' = J(y) with a measurable bounded positive function J. We meet also such situations when We solve the problem of the synthesis of optimal control. Here we move in an opposite direction.
358
CHAPTER 12
1. Sufficient conditions for equi-absolute continuity
Consider what role majorants of right hand sides of equations and inclusions play in connection with the question mentioned in the title of this section. The notion of equi-absolute continuity of sequences of functions (see §5.8) may be expanded in an obvious way on subsets of the space C.(M); namely, we say that a set Z ~ C.(M) is equi-absolutely continuous if: for every c > 0 there exists 8 > 0 such that if z E Z and a family {(ai, bi ) : i = 1, ... ,p} of pairwise disjoint intervals of the segment 1f( z) satisfies the condition I: {I bi - ai I: i = 1, ... , p} < 8 then I:{llz(bi ) - z(ai)11 : i = 1, ... ,p} < c.
Lemma 1.1. Let X, YI , ... , Y k ~ C.(IR x IRn). Let every function x E X may be represented as the sum of functions YI E YI ,.·. ,Yk E Y k : x = YI + ... + Yk· Let the sets YI , ... ,Yk be equi-absolutely continuous. Then the set X is equi-absolutely continuous too. Proof. In view of an obvious induction it is sufficient to prove the assertion for k = 2. I. Take an arbitrary E > 0 and fix 8 > 0 such that: if Y E YI U Y2, a family {( ai, bi ): i = 1, ... ,p} of pairwise disjoint intervals of the segment 1f(Y) satisfies the condition I:{lbi - ail: i = 1, ... ,p} < 8, then I)lly(b;) - y(ai)ll: i = 1, ... ,p} < E/2.
II. Let x E X and a family {(ai, bi ): i = 1, ... , p} of pairwise disjoint intervals of the segment 1f(x) satisfies the condition I:{lbi - ail: i = 1, ... ,p} < 8. The function x may be represented as the sum x = YI +Y2, where YI E YI and Y2 E Y 2· Since
Ilx(bi )
-
x(ai)11 = II(YI(bi ) - YI(ai)) + (Y2(bi) - Y2(ai))11 ~ IIYI(bi ) - YI(ai)11 + IIY2(b i ) - Y2(ai)ll,
then
I)llx(bi )
-
x(ai)ll: i
= 1, ... ,p}
~ I)IIYI(bi )
-
YI(ai)ll: i
+ L)IIY2(bi ) E E < - + - = E. 2 2 The lemma is proved.
= 1, ... ,p}
Y2(ai)1I : i = 1, ... ,p}
•
Complicated discontinuities in the space variables.
359
Lemma 1.2. Let -00 < a < b < 00. Let a function g : [a, bj - t ~ be generalized absolutely continuous, a function f3 : [a, bj - t ~n be measurable and 11f3(t)11 ~ -ftg(t) for almost all t E [a, bj. Then the function f3 is Lebesgue integrable. Proof. The function ~(t) = -ftg(t) is Denjoy integrable. From the inequality 11f3(t)11 ~ -ftg(t) it is non-negative. Theorem 4.9.1 implies its Lebesgue integrability. The function f3(t) is measurable. It is bounded in the norm by the Lebesgue integrable function Now Lemma 4.7.6 gives immediately what was required. The lemma is proved. • Assertion 1.1. Let U be subset of the product ~x~n, -00 < a < b < 00. For i = 1,2,... let hi : U - t [a, bj be a continuous mapping, a function O'.i : [a, bj - t [0,00) be measurable. Assume that:
e.
°
(1.1) for every c > there exists a number m E t E [a, b], O'.i(t) ~ m}) < c for every i = 1,2, ....
~
such that p,({ t
Let c > 0, {Zi: i = 1,2, ... } ~ Gs(U) be a sequence of continuous functions, the functions hi (t, Zi (t)) be generalized absolutely continuous, functions f3i : 7r(Zi) - t ~n be measurable and IIf3i(t)11 ~ c-fthi(t, Zi(t)), lIf3i(t) II ~ O'.i(hi(t, Zi(t))) for almost all t E 7r(Zi)· t Then the sequence of the functions Bi(t) = J f3i(s)ds, i = 1,2, ... , is equi-absolutely continuous. inf 7r(Z;)
Proof. Notice first that by Lemma 1.2 the functions f3i are Lebesgue integrable. Pass to main part of the proof. Take an arbitrary f > 0. According to (1.1) find m E [1,00) such that for every i = 1,2, ... the measure of the set
is less than c/2c. There exists an open subset Ei ~ Di of the segment [a, bj of the measure < c/2c. Let 8 = 2~ and i = 1,2, .... Let / = {(Pklqk): k = 1, ... ,s} be an arbitrary family of pairwise disjoint intervals of the segment 7r(Zi) and db) < 8. Let G k = {t: t E (Pk,qk), hi(t,Zi(t)) E Ed. The set G k is open and falls into the union of the at most countable family /k of its connected components.
360
CHAPTER 12
We have: qk
f
IIBi(qk) - Bi(pdll =
fJi(t)dt
Pk
f
=
fJi(t)dt
+L
{fJ fJi(t)dt:
J E
'Yk}
(Pk,qk)\G k
~ f
IlfJi(t)lldt + L
{fJ IlfJi(t)lldt:
J E
'Yk}
(Pk,qd\Gk
,;
m~((p" q,) \ G.) + L
~
m(qk - Pk)
C
{!
+ C L {hi(sup J, Zi(SUP J)) -
:t h,(t, z,(t))dt: J
E
~, }
hi(inf J, zi(inf J)): J E
'Yd.
Therefore
L {IIBi(qd -
Bi(pdll: k = 1, ... , s}
~ m2~ + C L {hi(sup J, Zi(SUP J)) -
hi(inf J, zi(inf J)) :
J E Ubk: k = 1, ... ,s}}.
Since {(hi(inf J, zi(inf J)), hi (sup J, Zi(SUP J))): J E Ubk: k = 1, ... ,s}} is a family of pairwise disjoint intervals lying in E i ,
L {hi (supJ, Zi(SUP J)) -
hi(inf J, zi(inf J)) :
J E Ubk: k = 1, ... ,s}} < JL(Ei ) ~ Thus
2)IIBi (qk) - Bi(pdll: k = 1, ... , s} <
E
E -.
2c
E
2' + 2' =
E.
The assertion is proved. • Example 1.1. Let U be an open subset of the product IF!; x IF!;n, K be a compact subset of the set U, C > 0. Let a function h : U ~ [a, b] ~ IF!; be continuously differentiable and a function a: [a, b] ~ [0, (0) be measurable. For (t,y) E U denote by H(t,y) the set
Complicated discontinuities in the space variables.
361
{u: u = (UI' ... ,Un) E lRn ,
Ilull : : ; c (
Oh oh Oh) at (t, y) + 0YI (t, Y)UI + ... + 0Yn (t, Y)U n ,
Ilull : : ; a(h(t, y))}. Assertion 1.1 implies immediately that the set D(H, K) is equi-absolutely continuous. The following assertion may be helpful when we verify the fulfilment of (1.1). Assertion 1.2. Let -00 < a < b < 00, i = 0,1,2, ... , functions ai : [a,bj ----t [0,00) be measurable and ai(t) ----t ao(t) for almost all t E (a, b). Then condition (1.1) holds. Proof. I. Since n{{ t: t E [a, b], ao(t) ~ m} : mE lR} = 0, there exists m E lR such that the measure of the set A = {t : t E [a, bj, ao(t) ~ m - I} is less than c, see Assertion 4.1.2. Thus the measure of the set B = {t: t E [a,b], ao(t) < m - I} is greater than b - a - c. By virtue of the convergence ai(t) ----t ao(t) almost everywhere, the measure of the set
BI = u{n{{t: t E B, ai(t) < m}:
i = i l ,i l +1, ... }: i l = 1,2, ... } ~ B
is equal to the measure of the set B. So for some i l = 1,2, ... the measure of the set n{{t: t E [a,b], ai(t) < m}: i = il,i l +1, ... }:;2 n{{t: t E B, ai(t) < m}: i = il,i l + I, ... } is greater than b - a-c. So the measure of the set u{ {t: t E [a, bj, ai(t) ~ m} : i = iI, i l + I, ... } is less than e. II. By virtue of arguments analogous to the first reasoning in I, there exists M > m such that for every i = 1, ... , i l the measure of the set {t: t E [a,bj, ai(t) ~ M} is less than c. III. By I and II for every i = 1,2,... the measure of the set {t : t E [a, b], ai(t) ~ M} is less than c. The assertion is proved. • As a consequence of Assertions 1.1-2 and Lemma 1.2 for n = 1, hi(t, y) = Y and f3i(t) = z;(t) we obtain: Assertion 1.3. Let -00 < a < b < 00 and U = lR x (a, b). For i = 0,1,2, ... let; a function ai : [a, bj ----t [0,00) be measurable, ai(t) ----t ao(t) for almost all t E (a, b),. functions Zi E Cs(U), i = 1,2, ... , be absolutely continuous and ~ z;(t) ~ ai(zi(t)) for almost all t E 7f(Zi)' Then the sequence {Zi : i = 1,2, ... } is equi-absolutely continuous. •
°
CHAPTER 12
362
Lemma 1.3. Let -00 < a < b < 00, -00 < c < d < 00, () > 0. Let a function g : [a, b] ---+ [c, dj be generalized absolutely continuous, a function a : [c, dj ---+ [0,00) be Lebesgue integrable, a function f3 : [a, bj ---+ ffi?n be measurable. Let ftg(t) ~ () and
11f3(t)11
(1.2)
~ a(g(t))
fOT almost all t E [a, bj. Then the function f3 is Lebesgue integrable. Proof. Under our assumptions the function g is increasing. Let A be an arbitrary primitive of the function a. By Remark 4.11.1 the function 1jJ(t) = A(g(z(t))) is absolutely continuous. By Corollary of Theorem 4.11.1 1
11f3(t)11
dId
~ a(g(t)) ~ (ja(g(t)) dtg(t) = (j dt 1jJ(t)
for almost all t E [a, bj. This implies what was required. The lemma is proved. • Assertion 1.4. Let K be a compact subset of the product ffi? x ffi?n, -00 < a < b < 00, (), 'fJ > 0. FOT i = 1,2, ... let hi : K ---+ [a, bj be a continuous mapping and a function ai : [a, b] ---+ [0,00) be Lebesgue integrable. Let functions Zi E Cs(K) and hi(t, Zi(t)) be generalized absolutely continuous. Let Ui E ffi?n and Iluill ---+ as i ---+ 00. Let functions f3i : 7r(Zi) ---+ ffi?n be measurable and
°
()
~
:t hi (t, Zi (t))
~ 'fJ + (Ui' z~ (t)),
(Ui' z~(t)) ~ 0, IIf3i(t)11 ~ ai(hi(t, Zi(t))) fOT almost all t E 7r(Zi)' Let the sequence of the functions t
Ai(t) =
J
ai(s)ds, i = 1,2, ... ,
a
be equi-absolutely continuous. Then the sequence of the functions t
Bi(t)
J
=
f3i(s)ds, i
=
1,2, ... ,
inf 71'(z;)
is equi-absolutely continuous.
Proof. Notice first that Lemma 1.3 implies the Lebesgue integrability of the functions f3i' i = 1,2, .... Pass to the main part of the proof. Take an arbitrary £ > O.
363
Complicated discontinuities in the space variables.
I. By virtue of the equi-absolute continuity of the sequence {Ai : i = 1,2, ... } there exists a number A > 0 such that
(1.3) for every family {( ak, bk ): k = 1, ... ,p} of pairwise disjoint intervals of the segment [a, b] with the total length ~ A and for every i = 1,2, ... we have
L{
1
Ci,(s)ds • k
=
~ 1 " p}
L {Ai(bd -
Ai(ad: k = 1, ... ,p} <
()E.
II. By virtue of the compactness of the set K and the convergence Ilui II -+ 0 there exists an index io = 1,2, ... such that (Ui' VI - V2) < ~ for every i = io + 1, io + 2, . .. and VI, V2 E K. III. By virtue of the absolute continuity of the functions B I , · · · , Bio there exists a number 80 > 0 such that if i = 1, ... , io and {( Ck, d k ) : k = 1, ... , p} is an arbitrary family of pairwise disjoint intervals of the segment 7r(Zi) with the total length ~ 80, then I:{IIBi(dd - Bi(cdll k = 1, ... ,p} < E. IV. Let 8 = min{ 80 , ~'2ArJ Show that if i = 1,2, ... and {(Ck' d k ) k = 1, ... , p} is an arbitrary family of pairwise disjoint intervals of the segment 7r(Zi) with the total length ~ 8, then I:{IIBi(d k ) - Bi(cdll i = 1, ... ,p} < E. For i = 1, ... ,io this follows from our choice in III. Let now i = io + l,io + 2, .... We have:
o < hi(d k , zi(d k ))
-
hi(Ckl zi(cd)
dk
=
J
:t hi(t, zi(t))dt
So the non-decreasing of the function (Ui' Zi(t)) (recall: (Ui' z~(t)) ? 0), the choice of io in II, and the inequality
L {d give the estimate: (1.4)
k -
Ck:
>.. k = 1, ... ,p} ~ 8 ~ 2TJ
CHAPTER 12
364 Now
dk
J
~
IIBi(dd - Bi(cdll
ai(hi(t, Zi(t))) dt
hi(dk,Zi(dk))
J
1 ()
ai(s)ds.
By (1.3-4) I:{IIBi(dd - Bi(cdll: k = 1, ... ,p} ~ E. The assertion is proved. • In order to explain the geometric contents of the just proved assertion word its particular case. When we take hi(t, y) = t and Ui == 0, we obtain Corollary. Let non-negative functions <po;, a E A, be Lebesgue integrable on the segment
[p, q],
and the family of the primitives t
Fo;(t)
=
J
<po;(s)ds, a E A,
p
be equi-absolutely continuous. Let Zo; denote the solution space of the inequality Ilyl(t)11 ~
Let us return to the fulfilment of (1.1). Assertion 1.5. Let real functions Ik, k = 1,2, ... , be defined on a segment [a, b], the sequence {Ik : k = 1,2, ... } be equi-absolutely continuous and adt) = f{(t) for almost all t E [a, b]. Then condition (1.1) holds. Proof. Assume the opposite. Then there exists a number E > 0 such that for every m = 1,2, ... the set Am = {k: p,({t: t E [a,b], adt) ~ m}) ~ E} is infinite. Since AI;2 A 2 ;2 A3;2 ... , we can fix indices k(l) < k(2) < k(3) < ... , k(i) E Ai. By Arzela's theorem we can make the choice of the indices k(i), i = 1,2, ... in a manner that the sequence {Jk(i): i = 1,2, ... } converges uniformly to an absolutely continuous function fo. Since p,({t: t E [a,b], ak(i)(t) ~ i}) ~ E we can fix a measurable set Mi ~ {t: t E [a, b], ak(i)(t) ~ i} with P,(Mi) = 7' By Lemma 5.9.1 the sequence {ak(i) : i = 1,2, ... } is weakly relatively compact. By Assertion 5.S.1
JCtk(i) Mi
(t)dt
~0
as
't
~
00.
365
Complicated discontinuities in the space variables. On the other hand, our choice implies
J
ak(i)(t)dt
~ ~i =
t:.
M.
The contradiction obtained gives what was required. • Example 1.2. Let: functions aij : IR ----t [0, (0), i = 1,2, j = 1,2, ... be locally Lebesgue integrable and their primitives be equi-absolutely continuous; functions Iij: IR x IR ----t [0,(0), i = 1,2, j = 1,2, ... be measurable in the first argument, be continuous in the second one and Iij(t,y) :S; aij(t) for all y E IR, i = 1,2, j = 1,2, ... ; K be a compact subset of the plane IR x IR; Zj P'l, A2) denote the solution space of the equation
and
Under these assumptions the set (ZO)K is equi-absolutely continuous. In order to prove it it is sufficient to show that every sequence {Zk k = 1,2, ... } <;;;; (ZO)K is equi-absolutely continuous. An usual passage to a subsequence reduces the question to the case when Zk E Zjk (A~, A~) and the sequences P7: k = 1,2, ... }, i = 1,2, converge (in the segment [0,1]). Let A? = lim A: and [ak' bk] = 1f(zd. k-HX!
Let h7(t, y) = A:t + (1 - Any. Put ;3;(t) = Iijk (A7t + (1 - AnZk(t), Zk(t)) for i Since the functions Iij are non-negative,
°
= 1,2 and k = 1,2, ....
:S; ;3;(t) :S; Iljk (A~t
+ (1
- A~)zdt), zdt))
+ I2jk (A~t + (1 -
A~)Zk(t), zdt))
= z~(t). Thus the functions ous). Since Zk(t)
;3; are Lebesgue integrable (their measurability is obvi-
= zk(ad + Bt(t) +
B~(t), where
B;(t) =
t
J ;3;(s)ds,
by
a.
Lemma 1.1 we obtain what was required when we prove that the sequence {B;: k = 1,2, ... }, i = 1,2, is equi-absolutely continuous.
366
CHAPTER 12
For A~ i- 1 we obtain what was required from Assertions 1.1 and 1.5 and the estimate
°
:::;;
k f3 i:::;; Zk t :::;; 1 _A~Ak + Zk t , I (
)
I (
)
=
d hki ( t, Zk ( t )) . 1 _1 Ak dt
,
For A~ = 1 we obtain what was required from Assertion 1.4 with TJ = 1, Uk = 1 - A~ (we consider here the scalar case) and when we take as f) an arbitrary number from (0,1) (the corresponding condition is fulfilled, beginning with some k). Example 1.3. Assume that TJ E [0,1); functions aij: ~ --t [0,00), i = 1,2, j = 1,2, ... , are measurable and for every c > there exists a number m E ~ such that for every i = 1,2, j = 1,2, ... JL({t: t E [a,b], aij(t)? m}) < C; functions iij : ~ x ~ --t [0,00), i = 1,2, j = 1,2, ... are measurable in the first argument, are continuous in the second and iij(t,y) :::;; aij(t) for all t, y E ~, i = 1,2, j = 1,2, ... ; K is a compact subset of the plane ~ x ~; Zj (AI, A2) denotes the solution space of the equation
°
and Under these assumptions the set (ZO)K is equi-absolutely continuous. This situation is analogous to that of Example 1.2 with small abbreviation of arguments because A~ i- 1 under the new assumptions. Example 1.4. Assume that functions aij : ~ --t [0,00), i = 1,2, j = 1,2, ... are measurable, functions alj : ~ --t [0,00), j = 1,2, ... , are locally Lebesgue integrable, and their primitives are equi-absolutely continuous, functions a2j : ~ --t [0,00), j = 1,2, ... , satisfy the condition:
°
for every c > there exists a number m E ~ such that JL{ t: t E [a, bj, a2j (t) ? m} < c for every j = 1, 2, .... Let functions iij : ~ X ~ --t [0,00), i = 1,2, j = 1,2, ... , be measurable in the first argument, be continuous in the second, and iij(t, y) :::;; aij(t) for all ~, i = 1,2, j = 1,2, .... Let K be a compact subset of the plane ~ x ~, () E [0,1), Zj(Al' A2) denote the solution space of the equation
Complicated discontinuities in the space variables.
367
and Under these assumptions the set (ZO)K is equi-absolutely continuous. This situation is analogous to these of two previous examples. Lemma 1.4. Let -00 < a < b < 00, -00 < c < d < 00, sequences of monotone functions fi : [a, b] ~ [c, d] and gi : [c, d] ~ ]Rn, i = 1,2, ... , be equi-absolutely continuous. Then the sequence of the functions gdi : [a, b] ~ ]Rn, i = 1,2, ... , is equi-absolutely continuous. Proof. The proof reduces to an obvious verification close to that of the proof of Assertion 1.4. • Example 1.5. Assume that U is an open subset of the product [a, b] x [c, d] x ]Rn, -00 < a < b < 00, -00 < c < d < 00, functions Cii, i3i, Ii : [c, d] ~ (0,00), i = 1,2, ... , are measurable, functions :: and
fi(U) =
j ::~:~
ds and if?i(U) =
a
j
a
be equi-absolutely continuous and the sequence of the functions i3i' i = 1,2, ... , satisfy (1.1). For t E [a, b] and x E [c, d]let
Fi(t, x) = {u: U E
]Rn,
Ilull ::;; li(x) +
Denote by Zi the solution space of the system
{
Xl E [ai(x),i3i(X) yl E Fi(t, x).
+
PutZ=U{Zi: i=1,2, ... }. Let K be a compact subset of the set U. Under these assumptions the set ZK is equi-absolutely continuous. Consider an arbitrary sequence {Zi: i = 1,2, ... } ~ Z K. Our aim is to prove that the sequence Zi = (Xi, Yi), i = 1,2, ... , is equi-absolutely continuous. The usual passage to a subsequence (with a corresponding renumbering) reduces the question to the case when Zi E Zi. Let [ai, bi ] = 7l'(zd. Represent the scalar function x; as the sum of two measurable functions Pi and qi: Pi(t) = max{O,x~(t) - .Bi(Xi(t))} and qi(t) = x~(t) - Pi(t) for t E [ai, bi ] . Then 0 ~ Pi(t) ~ 'Pi(t) and Qi(Xi(t)) ~ qi(t) ~ .Bi(Xi(t)) for almost all t E 7l'(zd·
368
CHAPTER 12
Likewise we represent the function functions ri and Si such that (1.5)
Ih(t)11
y;
~ li(Xi(t)) and
as the sum of two measurable
Ilsi(t)11
~
for almost all t E 1f( Zi). Lemma 1.1, Assertion 1.1 (for qi with hi(t, y) == y) and Corollary of Assertion 1.4 imply the equi-absolute continuity of the sequence of the functions Xi. The first inequality in (1.5) and the inequality x;(t) ~ ai(xi(t)) imply the estimate
Lemma 1.4 implies the equi-absolute continuity of the sequence of primitives of the functions rio The Corollary of Assertion 1.4 implies the equiabsolute continuity of the sequence of primitives of the functions Si. Now Lemma 1.1 implies the equi-absolute continuity of the sequence of the functions Yi. We have what was required.
2. Continuity of the dependence of solutions on the right hand side. First step Let U be an open subset of the product IR x IRn. In Assertion 5.9.1 we have obtained the possibility of proving that a sequence of spaces a = {Zi : i = 1,2, ... } <;;; Ri(U) and a space Zo E Ri(U) satisfy the condition:
(JJ) all functions from Ua are generalized absolutely continuous and if functions Zj E Zij (i 1 < i2 < ... ) are defined on a segment I <;;; IR and the sequence {Zj : j = 1,2, ... } converges uniformly to a function Z E Zo, then the function Z is generalized absolutely continuous and z' (t) E cc( {z; (t) : i = 1,2, ... }) for almost all tEl. Let us now discuss relationships and position of this property. Assertion 2.1. Assume that: (2.1) Fk : U ~ IR n , where k = 0,1,2, ... , are multi-valued mappings and the sequence {D(Fd: k = 1,2, ... } and the space D(Fo) satisfy condition (JJ) ;
°
(2.2) for every E > and k = 0,1,2, . .. there exists a set Ek <;;; U such that: 1,2,. .. and for every function Z E D(Fk ) the set a) for every k A = {t: t E 1f(z), (t, z(t)) f/. Ed is measurable and JL(A) < E,
Complicated discontinuities in the space variables.
369
b) for every point t E 1R the set (Eo)t contains the upper topological limit of the sequence of the sets {(Edt: k = 1,2, ... }, c) Fo(t,y) 2 n{cc(u{Fd{t} x V): k = i,i + I, ... }): i = 1,2, ... , V is a neighborhood of the point y in the set (Eo)d for every point (t, y) E Eo; (2.3) functions Zi E D(Fi) are defined on a segment I ~ 1R and the sequence {Zi: i = 1,2, ... } converges uniformly to a function Z E Cs(U). Then Z E D(Fo). Proof. When the segment I consists of one point only the assertion then is obvious because the space D(Fo) satisfies condition (p). Let the length d of the segment I is greater than zero and E E (0, d). Take sets Ek, k = 0,1,2, ... according to (2.2). For i = 1,2, ... the set Mi = {t: t E 7r(Zi), (t, Zi(t» rj. E i } is measurable and J.L(Mi ) < E. By the definition of a solution the measure of the set Ni of all t E 1\ M i , for which the approximate derivative z~(t) exists and belongs to Fi(t, Zi(t» is greater than d - E. The measure of the set Pi = U{Nk: k = i, i + 1, ... } (2 N i ) is greater than d - E. Since here PI 2 P2 2 ... , by Assertion 4.1.2 the measure of the set Q = n{Pi : i = 1,2, ... } is not less than d - E. Let Ql denote set of all t E Q satisfying the condition: the derivatives z'(t), z~(t), i
= 1,2, ...
exist and
Z'(t) E n{cc({z:{t): i ~ k}): k = 1,2, ... }. By virtue of (v) the measure of the set Ql is not less than d - E. Take an arbitrary t E Ql. Then t E Nk for an infinite number of indices k. That is, we can fix indices kl < k2 < ... such that t E N k · for j = 1,2, .... By (2.2,b) the point z{t) belongs to (Eok By (v) 1
= jo,jo + I, ... }): jo = 1,2, ... } V): k = i,i + I, ... })
Z'{t) E n{cc{{z~.{t): j 1 ~
CC{U{Fk{{t} x
for every i = 1,2, ... and for every neighborhood V of point z(t) in (Eo)t. By (2.2,c) z'{t) E Fo(t, z{t)). Let Qo denote the set of all points tEl, for which the derivative z'(t) exists and belongs to Fo(t,z(t». By the previous remarks J.L*(I \ Qo) < c. By virtue of the arbitrariness of E E (0, d) this implies that the set Qo has the full measure on the segment I. So z E D(F). The assertion is proved .• Theorem 2.1. Assume that:
CHAPTER 12
370
(2.4) F: U ---t lRn is a multi-valued mapping, D(F) E Rn(U) and Jor every c > there exists a set E ~ U such that: a) Jor every Junction Z E D(F) the set A = {t : t E 11"(z), (t, z(t)) ~ E} is measurable and J.L(A) < c, b) Jor every point t E lR the set E t is closed in Ut , c) F(t, y) :;2 n{ cc(F( {t} x V)) : V is a neighborhood oj the point y in Ed Jor every point (t, y) E E.
°
Then the set D(F) is closed in Cs(U). Proof. The proof consists in referring to Assertion 2.1 (with Fi == F) and to Lemma 6.4.2. Assertion 2.2. Assume that a = {Zi : i = 1,2, ... } E s(U), Z E Rp(U) and
•
(2.5) Jor every segment I oj the real line lR iJ a sequence oj Junctions Zj E Zij' 11"(Zj) = I, i1 < i2 < ... , converges to a Junction Z E Cs(U), then Z E Z. Then the sequence a converges to Z. Proof. The proof repeats the proof of Lemma 6.4.2. • Theorem 2.2. Let conditions (2.1), (2.2) hold. Let a = {D(Fi): i = 1,2, ... } E s(U). Then the sequence a converges to D(F). Proof. The proof consists in referring to Assertions 2.1 and 2.2. Corollary. Let condition (2.4) hold and D(F) ~ ~ E Rcnp(U). Then D(F) E Rc(U). By Theorem 4.13.4 we obtain the following Bokshtein theorem [Bo], [Faf] as another consequence of Theorem 2.2: Theorem 2.3 (Bokshtein). Let: 1) the right hand side oj a differential equation y' = J(t, y, a) depends continuously on the totality oj the variables y, a and is measurable in t; 2) a Junction cp(t) be integrable and IIJ(t, y, a)11 ~ cp(t) Jor all t, y, a. Then solutions oj the equation depend continuously on the parameter a (in the usual sense in the presence of the uniqueness of solution of the Cauchy problem and in the sense of convergence of sequences of solution spaces in the general case). • Example 2.1. Let measurable functions CPr : lR ---t [0, (0) depend on the parameter r E R ~ lR continuously in the following sense:
•
•
if ri ---t ro, then there exists a subsequence {rik: that CPr'k (s) ---t CPro (s) for almost all s E lR.
k = 1,2, ... } such
This happens, for instance, if functions CPr belong to the space L1 and the correspondence r ---t CPr is continuous with respect to the metric of the
Complicated discontinuities in the space variables.
371
space L 1, see Chapter 5, or if the function 'IjJ(r, s) = 'Pr(s) is continuous in each variable (apart). Let U be an open subset of the plane JR x JR, c > and a function f (t, y) from U into (c, (0) be continuous in the second argument, be measurable in the first, a function a: JR --t [0,(0) be integrable, and f(t,y) < a(t), 'Pr(t) < a(t) for every point (t, y) of the set U. Show that solutions of the equation y' = f (t, y) + 'Pr (y) depend continuously on the parameter r (in the sense of the convergence of sequences of solution spaces). Let r; --t ro and 'Pr, (s) --t 'Pro (s) for almost all t E lR. Let K be a compact subset of the set U, Z; E D(f + 'Pr" K), [a;, b;] = 7r(z;). For t E [a;, b;] put
°
(J;(t)
= f(t, z;(t)),
"I;(t)
= 'Pr, (z;(t)),
t
B;(t) =
J
(J;(s)ds,
t
fi(t) =
ai
J
"Ii(s)ds.
ai
The equi-absolute continuity of the sequences {Bi: i = 1,2, ... } and {fi: i = 1,2, ... } follows from Corollary of Assertion 1.4 and Assertion 1.1, respectively (the integrability of the functions "Ii follows from Lemma 1.2). By Lemma 1.1 the sequence of the functions Z; = zi(ai)+Bi(t)+fi(t) is equi-absolute continuous. This gives, in particular, the validity of the condition {D(f + 'Pr,} : i = 1,2, ... } E s(U). Let now "I = {(ai, bi ): i = 1,2, ... } be a family of pairwise disjoint intervals of the real line with db) ~ m. The condition f(t, y) ~ c implies that m JL( {t: t E 7r( z), z( t) E U"I}) ~ c for every solution z of every equation y' = f(t, y) + 'Pr, (y). Luzin's and Egorov's theorems, Theorem 2.2 and its Corollary imply that the solution spaces of the equation under consideration belong to Rc(U) and the sequence {D(f + 'PrJ : i = 1,2, ... } converges to the space D(f + 'Pro)' So we have proved the continuity of the mapping r --t D(f + 'Pr) E Rc(U), that means the continuity of the dependence of solutions of the equation y' = f(t, y) + 'P(r, y) on the parameter r. Example 2.2. Return to the notation of Example 1.2. Let (J" E (0,1). Let f1j(t, y) ~ (J" for all t, y E JR, j = 1,2, ... . 1. Let (a, b) <;; JR, i = 1,2, k = 1,2, ... , A = (h~)-1((a,b)) (= {(t, y) : t, y E JR, h~(t, y) E (a, b)}), z E Zoo Consider the function 'P(t) = h:(t, z(t)). Since
= .x: + (1
- .x:)z'(t) ~ .x~
+ (1 -
.x~)(J" ~
(J",
372
CHAPTER 12
the measure of the set {t: t E 7r(z), (t,z(t)) E A} is not greater than b~a. So if M is an open subset of the line, JL(M) :::;; (J and A = (hn-l(M), then
JL(({t: tE7r(Z), (t,z(t))EA}):::;; b~a. II. Let now >.~ - 7 >.~ as k - 7 00. Let flO, fzo : IR x IR -7 [0,00). For every fixed t E IR and for every segment I ~ IR let the sequence of the functions (Jij)t II converge uniformly to the function (JiO)t II" (Thus the functions fiO : IR x IR - 7 [0,00), i = 1,2, are measurable in the first argument and are continuous in the second). Show that the sequence of solution spaces of the equations
converges to the solution space of the equation
(2.7)
y'
= flO(>'~t + (1- >.ny,y) + f20(>'~t + (1- >'~)y,y) = Fo(t,y).
Let U be an arbitrary open subset of plane IR x IR with the compact closure. By the conditions imposed there exists a segment [c, d] of the real line IR containing all sets h{ ([U]), i = 1,2, j = 1,2, .... Let K be the projection of the compactum [U] in the y-axis. By Theorem 4.13.5 for every E > there exists a closed subset M of the segment [c, d] such that JL([c, d] \ M) :::;; ';, every function fij IMxK is continuous and the sequence {fij IMxK : j = 1,2, ... } converges uniformly
°
to the function fiO IMx K as j - 7 00. Let Ek = Un (n{(h{k)-l(M): i = 1,2, ... }). ByIJL({t: t E 7r(z), (t,z(t)) ~ Ej}:::;; E for every solutionz E D(Fk'U), By virtue of the convergence >.~ - 7 >.~ the set Eo is the upper limit of the sequence of the sets {Ek: k = 1,2, ... }. If Xk E Ek and Xk -7 Xo, then Fk(xd - 7 Fo(xo). Theorem 2.2 implies the convergence of the sequence of solution spaces of the equations (2.6) to the solution space of the equation (2.7) in the region U. Since the convergence of a sequence of solution spaces is a local property, we obtain the convergence of the sequence of solution spaces of the equations (2.6) to the solution space of the equation (2.7) on the entire domain IR x IR of the right hand side of the equation under consideration. Example 2.3. With the notation of Example 1.3 reasoning similar to that ones of Example 2.2 gives the following result. Let >.~ - 7 >.? as k -7 00. Let for every fixed value t E IR and for every segment I ~ IR the sequence of the functions (Jijk)t II converges uniformly to the function (fiO)t II" Then the sequence of solution spaces of the equations (2.6) converges to the solution space of the equation (2.7). We can continue the reasoning of Example 1.4 in an analogous way.
Complicated discontinuities in the space variables.
373
Example 2.4. Keep the assumptions of Example 1.5. Let for t E (a, b),
x E (c,d), y E JRn, when (t,x,y) E U, and for i = 0,1, ... the functions fli(X,y), hli(t,x,y) take values in JR, the functions fzi(X,y), h2i (t, x, y) take values in JRn, all they be measurable in the first argument and be continuous in the totality of the other ones,
:s; fli(X, y) :s; f3i(X), Ilfzi(X, y)11 :s; li(X), O:s; h1i(t, x, y) :s; !Pi(t), Ilh 2i (t, x, y)11 :s; !Pi(t).
Qi(X)
Assume that for a fixed value of the first argument and for every compactum K lying in the set of admissible values of other arguments we have the uniform convergence
flil{x}XK
--t
flOl{x}XK' f2il{x}XK
--t
f201{x}XK'
hlil{t}XK
--t
h10I{t}XK' h 2i l{t}XK
--t
h 20 1{t}XK'
When we repeat the arguments of the previous examples we obtain the convergence of the sequence of solution spaces of the systems
{
X' = fli(X,y) y' = f2i(X, y)
+ h1i(t,X,y), + h 2i (t, x, y)
to the solution space of the system
{ as i
X' = flO (x, y) y' = f20(X, y)
+ h10(t, x, y), + h 20 (t, x, y)
--t 00.
3. Existence theorems. First step According to the general remarks of §8.6 we obtain here existence theorems which correspond to the results of the previous section. Let us start with an example highlighting the general idea. It is trivial for our tools but it lies a very long way from the classical theory. Example 3.1. Let a function f : JR --t [1, (0) be measurable, and a function 9 : JR --t [0, (0) be locally Lebesgue integrable. Denote by Zj,g the solution space of the scalar differential equation (3.1)
y' = f(y)
+ g(t).
374
CHAPTER 12
In framework of the classical theory the complexity of the situation here consists in the possibility that its right hand side may have discontinuities both in t and in y simultaneously. Assume that (3.2) functions Ii : JR where,
--t
[1,00) are measurable and Ii
--t
I almost every-
(3.3) K is a compact subset of the plane JR x JR,
By Assertion 1.1 (with hi(t, y) = y and ai == Ii) and by Assertion 1.2 the sequence {Bi(t) = .h~fT((z;) Ii(Zi(S))ds: i = 1,2, ... } is equi-absolutely continuous. By Lemma 1.1 and results of §5.9 this implies the fulfilment of condition (v) for the sequence a = {Z/;,g: i = 1,2, ... }. By Theorem 2.2 and the Egorov theorem this implies the convergence (3.5)
Z/;,g
--t
Z/,g'
Notice that now we do not have the existence theorem and, respectively, we cannot be sure that the spaces Z/,g contain anything (besides solutions with one point domains of definition, which are always present). But if 'something' is in the spaces then 'it' converges. When we apply the assertion (about the convergence (3.5)) obtained to a stationary sequence Ii == I we obtain the fulfilment of the condition Z/,g E Rc(JR x JR), see Remark 7.6.1. Next, every function I of the mentioned type may be approximated by a sequence of continuous functions {Ii: i = 1,2, ... }, in the sense that Ii(t) --t I(t) as i --t 00 for almost all t, see Remark 5.13.1. Apply the Caratheodory's theorem to the equation y' = Ii(Y) + g(t) (see §8.2). So Z/"g E RcedJR x JR). By Theorem 8.1.1 the set Rcek(JR x JR) is closed in the topological space Rc(JR x JR). So the convergence (3.5) implies that Z/,g E Rcek(JR x JR). We have proved the theorem on the existence of solutions of the Cauchy problem for the equation (3.1) (and we have checked the fulfilment of the Kneser condition and the continuity of the dependence of solutions on initial values too). Return to Examples 1.4 and 2.3. We have to make a concrete construction. In order to be not restricted by the above understandings let us describe notation independently. Example 3.2. Let (3.6) a function ao : JR --t [0,00) be Lebesgue integrable, a function --t [0,00) be measurable, >. E (0,1), functions Ii : lR x JR --t [0,00),
al : lR
Complicated discontinuities in the space variables.
375
i = 0,1, be measurable in the first argument, be continuous in the second, h(t, y) ~ O!i(t) for i = 0,1 and for all t, y E R
Show that the solution space of the equation (3.7)
y' = fo(t, y)
+ f1 (At + (1
- A)y, y)
satisfies conditions (e) and (k). 1. Assume in addition that O!i == m > for every i = 0, 1. The Scorza-Dragoni theorem implies that we can fix closed subsets M1 ~ M2 ~ M3 ~ ... of the real line in a manner that JL(lR \ M j ) < 2-j and the functions fi IMj x[-j,jJ are continuous. Define functions fij in the
°
following way. Put fijlM' X [_'J,)'J = filM,} X[-' 'J' },} Let (c, d) be a connected component of the set lR \ M j • For t E (c, d) and y E lR we put: J
t-c fij(t, y) = d _ cfi(d, y)
d-t cfi(c, y)
+d_
(see Remark 5.13.1). We have established in Examples 1.4 and 2.3 that the sequence of solution spaces of equations (3.8)
y' = fOj(t, y)
+ f1j(At + (1
- A)y, y)
converges to the solution space of the equation (3.7) as j ----t 00. The right hand side of the equation (3.8) is continuous in the strip lR x (- j, j). Therefore its solution space there satisfies conditions (e) and (k). By Theorem 8.1.1 this implies that the solution space of the equation (3.7) satisfies conditions (e) and (k) too. II. Return to the consideration of the general case (that is, without the additional af:rmmption O!i == m > of I). For i = 0,1, ... ,p and j = 1,2, ... put fij(t, y) = min{j, fi(t, y)} and O!ij(t) = min{j, O!i(t)}. By I the solution spaces of the equations (3.8) (with the new meaning of Iij) satisfy conditions (e) and (k). The solution spaces of these equations Converge to the solution space of the equation (3.7) as j ----t 00. As in I this implies the fulfilment of conditions (e) and (k) for the solution space of the equation (3.7). In the general case remarks of Examples 1.4, 2.3 and 3.2 give: Theorem 3.1. Let condition (3.6) hold. Then the solution space of the equation (3.7) satisfies conditions (c), (e) and (k) and solutions of the • equation (3.7) depend continuously on A.
°
CHAPTER 12
376
An analogous construction based on the information of Example 2.4 gIves Theorem 3.2. Let:
-00 < a < b < 00, -00 < c < d < 00, U be an open subset of the product [a, bj x [c, dj x ~n; (3.9)
(3.10) functions a, (3" : [c, dj --+ (0,00) be measurable and functions
~
and
Let for t E (a, b), x E (c, d), y E ~n, when (t, x, y) E U, functions fl(x,y), hl(t,x,y) take values in~, functions f2(x, y), h 2(t, x, y) take values in ~n, be measurable in the first argument and be continuous in the totality of the others, a(x) ::::; fl (x, y) ::::; (3(x), Ilh(x, y)11 : : ; ,(x),
°: :;
hI (t, x, y) ::::;
y)11 : : ;
Then the solution space of the system (3.11)
{
x: = fl(x,y) y = f2(x, y)
+ hl(t,x,y), + h 2(t, x, y) •
satisfies conditions (c), (e) and (k).
Now generalize Theorem 3.2. We will obtain its version dealing with differential inclusions. Theorem 3.3. Let conditions (3.9 - 10) hold. For t E (a, b), x E (c, d), ~n, when (t,x,y) E U, let: Fl(x,y) and Hl(t,x,y) be (nonempty) segments (or one point subsets) of the real line, F2(x, y) and H 2(t, x, y) be (nonempty) convex compact subsets of the space ~n, mappings F l , F2, HI and H2 be measurable in the first argument and be upper semicontinuous in the totality of the others,
yE
Fl (x, y) ~ [a(x), (3(x)], HI (t, x, y) ~ [0,
lIull : : ; ,(x) lIull : : ;
for every for every
u E F 2 (x, y), u E H 2 (t, x, y).
Then the solution space of the system (3.12)
{
= Fl(x,y) y' = F2 (x, y)
X'
+ Hl(t,x,y) + H 2 (t, x, y)
Complicated discontinuities in the space variables.
377
satisfies conditions (c), (e) and (k). Proof. 1. The local fulfilment of condition (c) follows from the remarks of Example 1.5, of §4.13 and from the Corollary of Theorem 2.2. By the remarks of §8.3 this implies the global fulfilment of condition (c). II. Prove the fulfilment of conditions (e) and (k). Use the approximation of §8.2. In relation with our situation here the construction of §8.2 allows to state the existence offunctions !lj(X,y), !zj(x,y), h1j(t,x,y) and h2j (t, x, y), j = 1,2, ... (satisfying the conditions imposed in Theorem 3.2 on the functions !l(X,y), !2(X,y), h1(t,x,y) and h 2(t,x,y), respectively, with functions a, (3, , and cp of the statement of our theorem) and multivalued mappings Flj(x, y) 3 !lj(X, y), F2j (x, y) 3 !2j(X, y), Hlj(t,x,y) 3 h1j(t,x,y) and H 2j (t,x,y) 3 h 2j (t,x,y), j = 1,2, ... , satisfying the conditions imposed in Theorem 3.3 on the mappings Fl (x, y), F2(x,y), H1(t,x,y) and H 2(t,x,y), respectively, such that
(3.13)
Fll(x,y) ;2 F12 (X,y) ;2 F13 (X,y) ;2 ... , F21 (x, y) ;2 F22 (x, y) ;2 F23 (X, y) ;2 ... , Hl1 (t, x, y) ;2 H 12 (t, x, y) ;2 H 13 (t, x, y) ;2 ... , H 21 (t,X,y);2 H 22 (t,X,y) ;2 H 23 (t,X,y) ;2 ... , Fl (x, y) = n { Fli (x, y): i = 1, 2, ... }, F2 (x, y) = n { F2i (x, y): i = 1, 2, ... }, H1(t,x,y)=n{Hli(t,x,y): i=1,2, ... }, H 2(t, x, y) = n{H2i (t, x, y): i = 1,2, ... }.
By Theorem 3.2 the solution space of the system (3.14)
{
X' = !lj(X, y) + h1j(t, x, y), y' = !2j(X, y) + h2j(t, x, y)
satisfies conditions (c), (e) and (k). By I the solution space of the system (3.15)
{ X; = F1j(x, y) + H1j(t, x, y), y = F2j (x, y) + H2j(t, x, y)
satisfies condition (c). The solution space of the system (3.14) lies in the solution space of the system (3.15). By (3.13) and remarks of §8.2 the solution spaces of the systems (3.14) converge to the solution space of the system (3.12). By Theorems 3.2 and 8.1.1 the solution space of the system (3.12) satisfies conditions (c), (e) and (k). The theorem is proved. •
CHAPTER 12
378
4. Continuity of the dependence of solutions on the right hand side. Second step: complication of singularities Let:
(4.1) values of multi-valued mappings Fij : U --t JRn, i = 1, ... , q, j = 0,1,2, ... , of an open subset U of the product JR x JRn into the space JRn be nonempty, convex and compact
(4.2) single valued functions hp : U --t JR, p differentiable, m > 0, ~(t, y) ~ and
°
8h p( u} ~ t, ) Y uy}
=
1, ... , q, be continuously
8h p (t, Y) + ~ 8h p (t, Y) ~ + ... + Un ~ UYn ut
m
for all (t,y) E U, p = 1, ... ,q, j = 0,1, ... and u = (u}, ... ,un) E Fpj(t,y),
°
(4.3) < 'TJ < ~ and for every (t,y) E U, j and u E Fij(t, y)
= 0,1,2, ... ,
i,p
= 1, ... ,q
a) the angle between the vectors u and {~(t, y), ... , ¥v:(t, y)} be less than'TJ (or u =
0),
b)
p ( 8h 8y} (t, y)
p( ))2 )2 + ... + (8h 2 8Yn t, y ~ m .
Because of (4.3) (4.4) for every (t,y) E U and u,v E U{Fij(t,y) : j =-= 0,1,2, ... , i = 1, ... , q} the angle between the vectors u and v is less than 'TJo = 2'TJ < 7r (or one of them is equal to 0).
Let (4.5)
c = cos 'TJ
(> 0).
Under the assumptions of (4.1-3) consider the differential inclusion
y' E Gj(t, y),
j = 0,1,2, ....
Complicated discontinuities in the space variables. By (4.2) and (4.3,a) for every solution (4.6 j ) ah
z~(t)a P(t,z(t))
(4.7)
Yl
Z
=
ah
+ ... +«t)a
379
(Zl, ... ,zn) of the inclusion
P(t,z(t)) Yn
+
ah aP(t,z(t)) ~ m t
for almost all t E 7f( z). By Lemmas 4.9.6 and 10.2.1 all solutions of the differential inclusion (4.6 j ) belong to the set <1> of all functions z E Cs(U) satisfying the condition hp(t, z(t)) - hp(s, z(s)) ~ met - s) for every p = 1, ... , q and for every two points t > s of 7f(z) (if the domain of the function z consists of one point only then we put z E <1». See also §7.4. The verification of the fulfilment of the condition: (4.8) for every compactum K ~ U, for every index p = 1, ... , q, and for there exists a closed subset H of the compactum every number [ > hp(K) ~ IR such that J-L(hp(K) \ H) < [ and
°
Fpo(t,y) 2 n{cC(U{Fpk(h;l(H) i
= 1,2, ... , V
n V):
k = O,i,i
+ I, ... })
:
is a neighborhood of the point (t, y) in K}
is based on results of §4.13, see also the definition of the set E j in Example 2.2. Next, let: (4.9) functions
Q p :
IR
-+
[0,00), P = 1, ... , q, be locally Lebesgue integrable;
(4.10) Ilull ~ Qp(hp(t, y)) for every point (t, y) E U and every vector UEU{Fpj(t,y): j=0,1,2, ... }.
For p = 1, ... , q, M > 1 and for a generalized absolutely continuous function z E Cs(U) denote by 8~ (z) set of all generalized absolutely continuous functions y = (Yl, ... ,Yn) defined on the segment 7f( z) and satisfying the conditions:
(4.11) y(inf7f(z)) = 0; (4.12) Ilyl(t)11 ~ Mllz'(t)11 and Ily'(t)1I ~ Qp(hp(t, z(t))) for almost all E 7f(x).
t
380
CHAPTER 12
Assertion 4.1. Under assumptions (4.2 - 3) for every compactum K S;;; U and for every p = 1, ... , q the set b.~(K) = U{8~ (z) : z E U{D(Gj , K) : j = 0,1,2 ... } } is equi-absolutely continuous. Proof. By the first part of (4.12) and by the estimate
ddt hp(t, z(t)) =
z~ (t) ~hp (t, z(t)) + ... + z~(t) ~hp (t, z(t)) + 8: p (t, z(t)) uYn
UYI
ut
~ cmllz'(t)11
we have
lIy'(t)1I
~
M dd hp(t, z(t)). em t Assertion 1.1 implies immediately what was required (we apply it to the functions f3(t) = y'(t)). • Lemma 1.1 and Assertion 4.1 imply: Assertion 4.2. Under hypotheses of Assertion 4.1 the set
b.M(K) = b.~(K)
+ ... + b.~(K)
•
is equi-absolutely continuous.
Next: Lemma 4.1. Let M be a measurable subset of the real line. Let values
of measurable multi-valued mappings F, G : M ---t IR n be nonempty and compact. Let a measurable single valued mapping a : M ---t IR n satisfy the condition a(t) E F(t) + G(t) for almost all t E M. Then there exist measurable single valued mappings f3" : M that a(t) = f3(t) + ,(t), f3(t) E F(t), ,(t) E G(t) for almost all t EM. Proof. The condition a(t) E F(t)
+ G(t)
for almost all t E M
is equipotent to the condition
(a(t) - F(t)) n G(t) =J
0
for almost all t E M.
The measurability of the mapping
H(t) follows from Lemma 4.12.4.
= (a(t) - F(t)) n G(t)
---t
IRn such
Complicated discontinuities in the space variables.
381
Theorem 4.12.3 implies the existence of a measurable function "( : M ~ ffi.n, satisfying the condition
"(t) E H(t) for almost all t E M. The function
(3(t) = a(t) - "((t) is measurable,
a(t) - (3(t) = "((t) E a(t) - F(t). Therefore
(3(t) = a(t) - (a(t) - (3(t)) E F(t). The lemma is proved. • Lemma 4.2. Let M be a measurable subset of the real line. Let values of measurable multi-valued mappings Fi : M ~ ffi. n , i = 1, ... , q, be nonempty and compact. Let a measurable single valued mapping a : M ~ ffi.n satisfy the condition
a(t) E Fl (t)
+ ... + Fq(t)
for almost all t E M.
Then there exist measurable single valued mappings (3i : M ~ ffi. n, i = 1, ... , q such that a(t) = (31 (t) + ... + (3q(t), (3i(t) E Fi(t) for almost all t E M and for all i = 1, ... , q. Proof. For q = 1 put a == {31' For q = 2 we obtain what was required immediately from Lemma 4.1. Assume that q ~ 3 and assume that the assertion is proved for q - 1. The mapping is measurable. Its values are nonempty and compact. By Lemma 4.1 there exist measurable mappings {31, "( : M ~ ffi.n such that a(t) = (31(t) + "(t), (31(t) E Fl(t), "(t) E G(t) for almost all t E M. By the inductive hypothesis there exist measurable mappings (3i : M ~ ffi.n, i = 2, ... , q such that "(t) = (32(t) + ... + (3q(t), (3i(t) E Fi(t) for almost all t E M and for all i = 2, ... , q. This gives what was required. • Lemma 4.3. Under the assumptions of (4.1 - 2), (4.8) for every compactum K ~ U and for every number d > 0 there exists a closed subset C ~ K such that: 1) JL({t: t E 7r(z), (t,z(t)) rf; C}) :::;; d for every function z E
382
CHAPTER 12
Proof. According to (4.8) for E= d'; and for every p closed subset Hp of the compactum hp(K), such that
= 1, ... , q find
a
(4.13) and (4.14)
Fpo (t, y) :2 n { cc (U {Fpk (h; I (Hp) n V): k = 0, i, i
+ 1, ... })
:
i = 1,2, ... , V is a neighborhood of the point (t, y) in K}.
putc=n{h;I(Hp): p=l, ... ,q}nK. Let (t, y) E C. Let N be an arbitrary convex neighborhood of the compactum Go(t, y) in the space jRn. The mapping r
If)
••
TJ1)n
IN..
X ..• X
TJ1)n
IN..
---+ TJ1)n, IN..
1f)(V r I,···,
v) q = VI
+ . . . + V q,
is continuous. Therefore the set NI = rp-l (N) is a neighborhood of the compactum X = FlO(t, y) x ... x Fqo(t, y). It is not difficult to prove the existence of neighborhoods 0 FlO (t, y), ... ,0 Fqo (t, y) of the compactum FIO(t, y), ... , Fqo(t, y), such that OFIO(t, y) x ... x OFqo(t, y) ~ NI (use Lemma 2.4.2 and Theorem 2.1.2). By (4.14) and Lemma 2.5.2 for every index p = 1, ... , q there are a neighborhood Wp of the point (t, y) in the compactum K and an index i = 1,2, ... such that
For the set W
= WI n··· n Wq
and for every k
= i,i + 1, ...
we have
Gk(wnc) ~ N, U{ Gk(W n C): k = i, i + 1, ... } ~ N, cc(U{Gk(W n C): k = i,i + I, ... }) ~ N,
(4.15)
n{cc(u{Gk(wnc): k=i,i+1, ... }): i=1,2, ... , W is a neighborhood of the point (t, y) in K} ~ N.
In view of the arbitrariness in the choice of the neighborhood N of the compactum Go(t,y) (4.15) implies that
n{cc(u{Gk(wnc): k=i,i+1, ... }): i=1,2, ... , W is a neighborhood of the point (t, y) in K} ~ Go(t, y). This means the fulfilment of condition 2).
Complicated discontinuities in the space variables.
383
Let now Z E K. The measure of the set M = M1 U ... U Mq does not exceed ;; = d. This implies the fulfilment of condition 1). The lemma is proved. • When we consider a stationary sequence in which we repeat a mapping Fpj condition (4.8) goes into the condition:
(4.16 j ) for every compactum K ~ U, for every index p = 1, ... , q and for every number c > 0 there exists a closed subset H of the compactum hp(K) ~ lR such that f.L(hp(K) \ H) < c and
Fpj(t,y) :;2 n{cc(Fpj(h;l(H)
n V))
:
V is a neighborhood of the point (t, y) in K}. Lemma 4.4. Under the assumptions of (4.1 - 2), (4.16 j ), where j = 1,2, ... , for i = 1, ... , q and for every function Z E the mapping Fij(t, z(t)) is measurable. Proof. When we apply Lemma 4.3 to the stationary sequence FiO = Fi1 = ... = Fij = ... , we obtain from Theorem 2.5.2 that for every number d > 0 (and for every function z E 1. Do it. Let z E D(G j , K). By Lemma 4.4 the mappings
F1j(t, z(t)), ... ,Fqj(t, z(t)) are measurable. By Lemma 4.2 there exist functions /31, ... ,/3q : 7r(z) -+ lRn such that z'{t) = /31 (t)+· ·+/3q(t) and /3 p(t) E Fpj(t, z(t)) for all i = 1, ... ,q and for almost all t E 7r( z). It remains to show that there exists a number M > 0 (which does not depend on p, j and z) such that the function t
Yi(t)
=
J /3i(S) inf 7r(z)
ds
384
CHAPTER 12
belongs to t5{'1(z). (The fulfilment of (4.11) is obvious.) Let p = 1, ... ,q, t E 7r(z),
By (4.4) the angle between the vectors u and v is less than 'TIo. Therefore
where d = cOS'TIo
i=
±1,
Ilu + vll 2> (liull + dllvll)2 + (1 - d2)IIull 2 ~ IIu + vII > IIu II Vl-=d2. Thus
11,8 (t)II < II,81(t) p
(1 -
d2)IIull 2,
+ ... + ,8q(t) II < IIz'(t) II
JI=d2
JI=d2
and the first part of condition (4.12) holds with M = v'1~d2. The fulfilment of the second part of (4.12) follows immediately from (4.10). The assertion is proved. • Assertion 4.3, Lemma 4.3 and Theorem 2.2 imply: Assertion 4.4. Under the assumptions of (4.1 - 6), (4.8 - 10) and (4.16 j ), j = 1,2, ... , the sequence of solution spaces {D(G j ) : j = 1,2, ... } converges to D(G o ).
•
5. Existence theorem. Second step
Assume that: (5.1) values of multi-valued mappings FiO : U ---+ JRn, i = 1, ... , q, of an open subset U of the product JR x JRn into the space JRn are nonempty, convex and compact, Go(t, y) = F10(t, y) + ... + Fqo(t, y); (5.2) single valued functions hp : U ---+ JR, p differentiable m > 0, 7lf(t, y) ~ 0 and Ul
ohp
~(t,y) UYI
for all (t,y)
E
(5.3) 0 < 'TI <
=
1, ... , q, are continuously
ohp ( ) ohp ( ) + ... + un~ t,y + "'t t,y ~ m UYn
U
U, p = 1, ... ,q, and u = (Ul, ... ,un) ~
E
Fpo(t,y);
and for every (t, y) E U, i,p = 1, ... ,q and u E FiO(t, y)
Complicated discontinuities in the space variables.
a) the angle between the vectors u and 1/ (or u
{F:;(t, y), ... , ~(t, y)}
385
is less than
= 0),
b) ( )) 2 ( 8hp 8Yl t, Y
p
( )) 2 + ... + (8h 8Yn t, y
2
~ m ;
(5.4) condition (4.9) holds, Ilull ~ CY.p(hp(t, V)) for every point (t, y) E U and every vector u E Fpo(t,y), condition (4.16 0 ) holds. Assertion 4.4 implies: Assertion 5.1. Under the assumptions of (5.1- 4) the space D(G o) of solutions of the differential inclusion y' E Go(t, y)
(5.5)
belongs to Rc(U), • Our next target is to prove the existence theorem for differential inclusion (5.5). Following our general ideas we will construct a 'good' approximation of the right hand side. We need not construct such an approximation for the entire region U all at once and it is sufficient to construct the needed approximation for a neighborhood of its arbitrary point. Properties in question have local character. Assertion 5.2. Under the assumptions of (5.1-4) for every point x E U there exists its neighborhood V ~ U, such that D(G o, V) E [Rceu(V)].
Proof. A. Prove first our assertion with the additional assumption that the functions CY.i are constant. 1. Fix a number c E (0,1) such that [Ocx] ~ U and for every point (t, y) E [0 c x] and every i = 1, ... , q the angle between the vectors
p 8h p } { 8h 8Yl (t,y), ... , 8Yn (t,y) and ei
8hp ( 8h p } = { 8Yl x), ... , 8Yn (x)
is less than 1/1 = H~ - 1/). This is possible by virtue of the continuity of the functions ~. By (5.3):
386
CHAPTER 12
(5.6) for every (t,y), (to,yo) E [Oex] and angle between the vectors u and
U
E
U{Fi(t,y): i = 1, ... } the
( ) 8h p } { 8hp 8Y1 to, Yo , ... , 8Yn (to, Yo)
H
is less than "12 = "1 + 2"11 = "1 + ~ - ¥ = ~ + "1) < ~. Fix an arbitrary number r E (0, c:) and put V = Orx. II. Let i = 1, ... ,q. Put fi = grad hi(x). Let Xo E V and
:!.-hi(
? lIei 11m cos "11 ? > O.
+ ... + 8h i (x) 8h i (
8Yn 8Yn 8h i 8h i + -(x)-(
8t
8t
m2 cos "11
So: (5.7)
if a
~
a' < {3'
~
(3 then
III. With the notation of II let
o < 00 <
m 2 (c: - r)cOS"11
Ilf;!1
'
0 E (0,1)
and the measure of an open subset A of the real line IR not exceed 00 0 . Then the measure of every connected component A1 of the set A does not exceed 00 0 too. By (5.7) and by the definition of 00 :
Hence
X2
E
O,x.
Complicated discontinuities in the space variables.
387
So if a connected component of the set 1(xo, ei) n hi 1 (A) meets the set V then its lengt h is less than (} (E - r). So it lies entirely in 0 EX. IV. With the notation of III associate to every point s E hiI(Ad n V the endpoints bI(s) and b2 (s) of the intervall(s,fi)nhil(Ad (choose the point bi (s) and b2 (s) in a manner that the vector b2 (s) - bi (s) is co-directed with fi). This choice of bi (s) and b2 (s) does not depend on the choice of Xo. The functions bl , b2 : hil(Ad -+ ~n+l are continuous. This follows from our choice in I. V. In IV we have defined values of bi and b2 for all connected components of the set A. Define a multi-valued mapping
Ilbl (x) - xii _ Ilb 2 (x) - bl (x)II Fi (b 2 (x)) Ilb 2 (x) - xii _ _ bi (x)11 Fi(b i (x))
+ Ilb 2 (x)
if hi(x) E A.
In this definition
IlbI(x) - xii Ilb 2 (x) - xii - 1 Ilb2 (x) - bI(x)11 + Ilb 2 (x) - bl(x)11 - . Therefore
FiA(X) ~ CC(Fi(bl(x)) U Fi(b 2 (x))). Now select a set A such that the mapping
Fi Ih;-l(IR\A)nV
is upper semi-
continuous. Then the mapping FiA is upper semicontinuous too, see Remark 4.13.l. VI. The mapping GA = FIA + ... + FqA is upper semicontinuous. Therefore D(G A , V) E [Rceu(V)]Rc(V), see §8.2. VII. By V and (5.6) for every point (t, y) E V and for every vector
U = (UI, ... ,u q)
E
U{FiA(t,y): i = 1, ... ,q}
the angle between the vectors U and ei is less than 772. VIII. Fix Bo according to III. Fix a sequence 1 > (}I > B2 > ... , Bj -+ 0 and open subsets Al 2 A2 2 ... of the real line such that Jl(A;) ::;:; BiBO and the mapping Fplh;l(IR\Ail is upper semicontinuous. Return to the considerations of the previous section and put By Assertion 4.4 D(Gi,V) -+ D(G o, V).
Fpi
=
FpAi .
388
CHAPTER 12
By VI this gives what was required. Under the additional assumption A the assertion is proved. B. Pass to the general case Put {3 ·(t) = J PJ max{j, ap(t)} for p = 1, ... ,q, j = 1,2, ... and t E JR. Put Fpj(t, y) = {3pj(h p(t, y))Fp(t, y) for p (t,y) E U. Evidently
Ilull :::; j
=
1, ... , q, m
=
1,2, ... and
for every vector u E Fpj(t, y)
and the mappings Fpj satisfy all conditions of the previous section. By Assertion 4.4 D(G j ) -+ D(G o) as j -+ 00. Because of A this implies that D(G o, V) E [Rceu(V)]. The assertion is proved. • By Theorem 8.1.1 Assertions 5.1 and 5.2 imply: Theorem 5.1. Under the assumptions of (5.1 - 4) the solution space of the differential inclusion (5.5) belongs to Rcek(U). • In other words for differential inclusion (5.5) assertions about the existence of solutions of the Cauchy problem and about the continuity (the upper semicontinuity) of the dependence of solutions on initial values are true. Example 5.1. Theorem 5.1 is applicable in an obvious way to the scalar equation (5.8)
where the functions fl' f2 : JR -+ [0,(0) are Lebesgue integrable (in order to analyze the situation we may consider the functions hI (t, y) = t + y + y3 and h 2(t, y) = t + 2y + siny). Assertion 4.4 implies, for instance, the following assertion about the continuity of the dependence of solutions of the equation (5.8) on the right hand side: if functions cp : JR -+ [0,(0) are locally Lebesgue integrable, real functions gl(t,a) and g2(t,a) are non-negative and measurable in t, are continuous in a, and Ig1 (t, a)1 ~ cp(t), Ig2(t, a)1 ~ cp(t), then solutions of the equation y' = gl (t
+ y + y3, a) + g2 (t + 2y + sin y, a)
depend continuously on the parameter a.
CHAPTER 13
EQUATIONS AND INCLUSIONS OF SECOND ORDER. CAUCHY PROBLEM THEORY
Our aim in this chapter is to see what specific character an increment in the order of equations and inclusions under consideration brings to the structure of the theory of the Cauchy problem, and how this may imply the possibility of a weakening of restrictions on functions appearing in equations (inclusions) . Here we will restrict our considerations to the scalar case. 1. Cauchy problem theory for differential inclusions of second order
According to the general remarks of §6.1 we regard a differential inclusion (1.1)
x"
E
F{t,x,X')
Xi
=y,
as equivalent to the system (1.2)
{
yl E F(t, x, y),
or to the vector inclusion (1.3)
u l E Fo(t, u),
where
Let: (1.4) -00 ~ a < b ~ 00, -00 ~ c < d ~ F : U - 7 IR be a multi-valued mapping. Outside the set (1.5)
E
00,
U
= (a, b) x «c, d) x {O})
=
(a,b) x «c,d) x IR),
CHAPTER 13
390
methods of previous chapters are applicable without any essential modification to the analysis of the situation (see, for instance, Theorem 12.3.3), and for some time we leave this question apart. In order to analyze the structure of the solution space of the equation (1.3) near the set E introduce the estimates (1.6) a(t,x) = liminf (inf F(t,x,y)) and {J(t, x) = limsup(supF(t,x,y)). u---+x,s-+t
u---+x s---+t
y-->O
y--':'o
We will assume that (1.7)
if
0 E [a(t, x), (J(t, x)],
then
0 E F(t, x, 0).
Condition (1.7) turns out to be sufficient to extend the fulfilment of condition (c) from the set (1.8)
V=U\E
to the entire region U. Since condition (c) is a particular case of the general notion of the convergence of sequences of solution spaces, it is more convenient to analyse the general situation. Let (1.9) F;*: U
-t
~,
i = 1,2, ... , be a sequence of multi-valued mappings.
For u =
(~)
put (1.10)
Fi(t, u)
=
(F*(/ i ,x, Y)) .
Assertion 1.1. Let conditions (1.4 - 5), (1.8 - 9) hold, (1.11) ao(t, x) = liminf (inf F;*(t, x, y)) and (Jo(t, x) = lim sup (sup F;*(t, x, y)), u--+x :s---+t y---+O,t--+oo
(1.12)
if
0 E [ao(t, x), (Jo(t, x)],
U--1oX ,s--+t y--+O,i---+oo
then
0 E F(t, x, 0),
(1.13) with the notation of(1.3), (1.10) the sequence of the spaces {D(Fi , V) : i = 1,2, ... } converge in V to D(Fo, V). Then
Equations and inclusions of second order. Cauchy problem theory.
(1.14) the sequence {D(F;): D(Fo).
i
=
391
1,2, ... } converges in U to the space
Proof. I. We start with the discussion of the fulfilment of the condition {D(Fi) : i = 1,2, ... } E s(U). Let K be an arbitrary compact subset of U. In the region V we have the needed convergence (see (1.13». So it is sufficient to analyze the set of singularities E = U \ V. Use Assertion 7.3.2. Put such that K ~ (a, b) x (c, d) x [-m, m]. Then for z E (Zj)K, j = 1,2, ... , the function
°
lim top sup D(Fi)
~
D(Fo).
i-+oo
The set
Eo = {(t,x,O): (t,x,O) E E,
°
E
[ao(t,x),,Bo(t, x)]}
is closed. Consider an arbitrary point (tl' Xl, 0) E E \ Eo. By our definition of the set Eo there are a number E E (0, min{ tl - a, b - td) and an index i such that either:
U{F;(t,x,y): k=i,i+1, ... , (t,x,Y)EW(thXd}~[E,oo), where W(tl, Xl) = (tl - E, tl + E) X (Xl - E, Xl + E) X (-E, E); or:
U{F;(t, X, y): k
= i, i + 1, ... , (t, X, y)
E W(tl, xd} ~ (-00, -E).
Both cases are analogous. It is sufficient to restrict our consideration to the first one. The function h : V(tl' xd - t JR, h(t, X, y) = y, satisfies the condition
h(t,z(t»-h(s,z(s» ~ c(t-s) for every functionz E U{D(Fk' W(h,XI» : k = 1,2, ... } (by virtue of the estimate y'(t) ~ E, where z = (x, y». The remarks of §7.5 (or, for instance, Theorem 7.6.2) imply the convergence D(Fk' W(tl' xd) - t D(Fo, W(tl, xd).
In order to complete the proof consider an arbitrary function z = (x,y) E lim top sup D(Fk)' k-+oo
392
CHAPTER 13
Let
M = {t: t E 1r(z), (t, z(t)) E Eo}. By Theorem 7.6.4 (with
U,
=
Vo)
lim top sup D(Fk' Vo)
<,;:;
D(Fo, Vo),
k->oo
where
Therefore: if a segment I lies in the set 1r(z) \ M, then
ZiI
E D(Fo, Vo)
<,;:;
D(Fo).
It remains to check that:
z'(t) E Fo(t, z(t)) for almost all t E M.
(1.15)
yiM
Notice that
== O. Therefore:
y'(t) = 0 for almost all t E M.
(1.16)
In addition, when we pass to the limit in the equality t
Xi(t) = xi(inf1r(zi))
+
J
Yi(s)ds
inf 7f{Z;}
we obtain that
t
x(t) = x(inf1r(z))
+
J
y(s)ds.
inf7f(z)
This gives the fulfilment of the first equation of (1.1). This fact and (1.16) (see the definition of Eo) mean the validity of (1.15). Thus z E D(Fo) what was required. The assertion is proved. • When we put Ft == F, we obtain: Corollary. Let conditions (1.4 - 8) hold and D(Fo, V) E Rc(V). Then
D(Fo)
E
Rc(U),
•
Assertion 1.2. Let conditions (1.4 - 8) hold and
D(Fo, V)
(1.17)
Then D(Fo)
E
Rce(U).
E
Rce(V).
Equations and inclusions of second order. Cauchy problem theory.
393
Proof. By the Corollary of Assertion 1.1 it remains to prove that the space Z = D(Fo) satisfies condition (e). Do it. The set Eo = {(t,x,O): (t,x,O) E E, E [a(t, x),{J(t, x)]} is closed. I. If (tl,ud E V, then (1.17) implies what was required. II. If a point (tl,Ul) = (tl,Xl,O) belongs to the set E \ Eo, then we repeat the reasoning of the step II of the proof of Assertion 1.1. Fix a neighborhood W(tl,xd of the point (tl,Xl), the function h(t,x,y) = y and a number c > such that
°
°
h(t, z(t)) - h(s, z(s))
~
c(t - s)
for every function z E D(Fo, W(tl,Xl)). N ow the fulfilment condition (e) for the point (tl' ud follows from Theorem 11.4.1. III. To complete the proof consider the family of all open rectangles ~, which lie in the plane ~x (~x {O}), have sides parallel to the coordinate axis, have tops with rational coordinates, and satisfy the condition [~l ~ E\Eo. The family S is countable: S = {~i: i = 1,2, ... }. Define the mapping F;* : U ---t ~:
F,*(t, x, y) =
1
F(t, x, y)
cc(F(t, x, y) U {O})
°
if y =J or if (t,X,y)E~lU···U~i' in the opposite case.
The mapping F;* satisfies all conditions imposed on the mapping F III Corollary of Assertion 1.1. So D(Fi) E Rc(U) (see notations in (1.10)). Here F(t, x, y) ~ F;*(t, x, y) for all (t, x, y) E U. Therefore by I and II the solution space D (Fi) satisfies condition (e) for (t, u) E U \ Eo. If (t, u) E Eo, then (t,u) tj. [~l U··· U ~il. Therefore for some 'TJ E (O,min{t - a, b - t}) the segment [t - 'TJ, t + 'TJl x {u} lies outside ~l U··· U ~i. This means that the function z = u defined on the segment [t - 'TJ, t + 'TJl belongs to D(Fi). Thus D(Fi) E Rce(U). Since F(t, x, y) = n{F;*(t, x, y): i = 1,2, ... } for every (t, x, y) E U, remarks of §8.2 imply the membership D(Fo) E Rce(U). The assertion is proved. • Together with above methods allowing to analyze the behavior of solution spaces on the set V new remarks give a sufficiently full set of tools to prove the fulfilment of basic properties of solution spaces. We will not try to give just now a complete picture and we will restrict further consideration by a case, which is particular for framework of the general theory. Assume that the right hand side F may be represented as the sum of two mappings (1.18)
F(t,x,y) = G(x,y)
+ H(t,x,y)
394
CHAPTER 13
satisfying slightly weakened Davy conditions, namely: (1.19) values of the mappings G and H are segment and one point subsets of the real line IR; (1.20) for every fixed x E (c, d) and y =I- 0 the mapping Gx(Y) = G(x, y) is upper semicontinuous; (1.21) for every y =I- 0 the mapping GY(x) = G(x, y) is measurable; (1.22) for every compactum K ~ (c, d) x (IR \ {O}) there exists a integrable function cp : (c, d) ---t [0,00) such that G(x, y) ~ [-cp(x), cp(x)] for all (x,y) E K; (1.23) for every fixed t E (a, b) and y =I- 0 the mapping Ht(x, y) is upper semicontinuous; (1.24) for every y measurable;
f. 0 and x
E
= H(t, x, y)
(c,d) the mapping HX,Y(t) = H(t,x,y) is
(1.25) for every compactum K ~ V there exists an integrable function cp: (a,b) ---t [0,00) such that H(t,x,y) ~ [-cp(t),cp(t)] for all (t,x,y) E K. Theorem 12.3.5 and Assertion 1.2 imply: Theorem 1.1. Let conditions (1.4), (1.18-25) and (1.6-7) hold. Then • the solution space of the system (1.2) belongs to Ree (U).
2. Relatively weakly compact families of majorants The increment of the order of an equation (inclusion) brings particularities in the consideration of the question about the continuity of the dependence of solutions on the right hand side. In this section we will point out one of such new relations. Let: (2.1) -00 ~ a < b ~ 00, -00 ~ c < d ~ 00, U
{
x'(t) = y(t), y'(t) E [CP1(t)
= (a,b)
x ((c, d) x IR).
+ 'ljJl(X(t», CP2(t) + 'ljJ2(X(t»].
Start with a preliminary remark.
Equations and inclusions of second order. Cauchy problem theory.
395
Lemma 2.1. Let with the notation of (2.1) (2.3) Xi: [a, b] -+ (e, d), i = 1,2, ... , be a sequence of absolutely continuous non-decreasing functions, the sequence {x~ : i = 1,2, ... } be relatively weakly compact in L 1([a,b]),
(2.4) pact,
a sequence {'l/Ji: i = 1,2, ... } <;;; L 1([c,d]) be relatively weakly comt
Wi(t) = J'l/Ji(s)dS. c
Then the sequence of functions {y~: i = 1,2, ... }, where Yi(t) = Wi(Xi(t)), is relatively weakly compact. Before to start the proof notice that by Remark 4.11.1 and Assertion 4.11.1 the functions Wi(Xi(t)) are absolutely continuous and
(2.5)
(Wi(Xi(t)))' = 'l/Ji(Xi(t))X:(t)
t E [a, b].
for almost all
Proof. By Remarks 5.8.1 and 5.8.2 and Theorem 5.8.1 we obtain what was required from Lemma 12.1.4. • Assertion 2.1. Let with the notation of (2.1 - 2):
(2.6) the sequences {
Z
= 1,2, ... } <;;; L1 ([e, d]),
j
1,2, converge
(2.8) the sequence of the mappings
and the mapping F(t, X, y) = [ be the lower and M ? m be the upper bounds of the projection of the compactum K in the third factor of the product (a, b) x ((c, d) x (0,00)). Let Zj = (Xj, Yj) E (Zj)K, where the index j runs Over an infinite subset C of N, 7r(Zj) = [aj, bj ].
°
396
CHAPTER 13
Our aim is to prove that the sequence {Zj : j E C} has a subsequence converging to a function Z E Z. Do it. II. Since the sequences of the primitives of the functions in (2.6) and (2.7) are equicontinuous, without loss of generality we can assume in addition that the sequences {
Pi(t)
=
if y;(t)
~
and qi(t) = y;(t) - Pi(t). Here:
(2.9) (2.10)
Pi(t) E [
As a consequence of (2.10) we obtain:
Iqi(t)1
(2.11)
~
'l/J;(t)
for almost all t E [ai, bi ],
where 'l/J:(t) = l'l/Jli(Xi(t)) I + 1'l/J2i(Xi(t))I· By the first equation of (2.2) (2.12) 0
<m
~
x;(t)
~
M for all i E C and for almost all t E [ai, b;].
Therefore (2.11) implies
1 1 (2.13) Iqi(t)1 ~ -'l/J:(Xi(t)) . x;(t) = -(w;(x(t)))' for almost all t E [ai, bi ], m m where u w;(u) =
J
'l/J:(s) ds.
a
Equations and inclusions of second order. Cauchy problem theory.
397
The sequence of the functions {w;: i = 1,2, ... } is equi-absolutely continuous, see Remark 5.8.4 and §5.9. Now extend the function Xi to the entire segment [a, b], by putting Xi(t) = xi(ai) for t :::; ai and Xi(t) = xi(b i ) for t ~ bi . The extended function Xi remains a solution of the inequality 0 :::; x;(t) :::; M. Therefore the sequence of the extended functions {Xi: i E C} is relatively weakly compact. By Lemma 2.1 the sequence of functions {(W;(Xi(t)))' : i E C} is also relatively weakly compact. Now (2.13) implies the relative weak compactness of the sequence of the functions {qi: i E C} (in order to have a possibility of correct references extend the function qi by zero on [a, b] \ [ai, biD. The relative weak compactness of the sequence of (extended in an analogous way) functions Pi, i E C, follows from (2.9). These facts imply the relative weak compactness of the sequence of (extended in an analogous way) functions y~, i E C. By just made remarks we can regard without loss of generality that the sequences {Pi: i E C} and {qi: i E C} converge weakly to functions P and q, respectively. Now our remarks and (2.12) imply the equi-absolute continuity of the sequence {Zi: i E C}. Therefore without loss of generality we can regard that the sequence of the functions {Zi: i E C} converge uniformly to an absolutely continuous function Zo = (xo, Yo). For lao, bo] = 7f(ZO)
ao = lim ai, bo = lim bi iEC iEC
(2.14)
i-+oo
i-+oo
Take arbitrary points a < f3 of the interval (ao, bo). By (2.14) [a,f3] ~ (ai,b i ), beginning with some i = io E C. For t E [a, f3] we have: t
xo(t) =
~~~ i--+oo
Xi(t) =
~~~
xi(a)
+ ~~~
i--+oo
i-+oo
t
J
Yi(S) ds = xo(a)
+
Therefore: (2.15)
X~(t) =
Yo(t)
for all
t E
[a, f3].
Next, for to E (a, t) by (2.9) we have: t
4>li(t) - 4>li(to) :::;
J
Pi(S) ds :::; 4>2i(t) - 4>2i(t O)'
to
J
Yo(s) ds.
0:
Q:
398
CHAPTER 13
When we pass in these estimates to the limit as i we obtain
~ 00,
i E C, i ;:: i o,
t
J
~
p(s) ds
to
~
J () t
~
1 -t - to
p s
~
t - to
to
~
The passage to the limit as t Thus:
ds
.
to is possible for almost all to E (a, (3).
(2.16) 'PI(tO) ~ p(t o) ~ 'P2(t O) for almost all to E (a,(3).
Next, for to E (a, t) by (2.10) we have:
Wli(Xi(t)) - Wli(Xi(t O))
~
t
J J
qi(S)Yi(S) ds
to
t
=
t
qi(S)(Yi(S) - Yo(s)) ds
to
+
J
qi(S)YO(s) ds
to
When we pass here to the limit as i t
~ 00,
i E C, i ;:: io the estimates
t
J
JIqi(S)I·IIYi - Yoll ~ IIYi - Yoll JIqi(S)1 ~
qi(S)(Yi(S) - Yo(s)) ds
~
ds
~
t
ds
~ 0,
to
imply that:
J J t
wI(xo(t)) - wI(xo(t O))
~
q(s)yo(s) ds
~
w2(xo(t)) - w2(xo(tO)),
to
t
wI(xo(t)) - wI(xo(tO) ~ _1_ t - to '" t - to
(s) (s) ds ~ w2(xo(t)) - W2(XO(to)). q Yo '" t - to
to
So as in (2.16)
'ljJl(XO(tO))Yo(to) ~ q(to)Yo(t o) :::; 'ljJ2(XO(tO))Yo(to)
Equations and inclusions of second order. Cauchy problem theory.
399
or:
Since t
t
y~(S)
Yo(t) - Yo(a) = j
.!~rg
ds =
0:
'l--+OO
j y;(s) ds Q
t
= lim iEC
i--+ 00
t
jPi(S) ds
t
= j p(s) ds <>
+
lim jqi(S) ds iEC
i--+ 00
Q
Q
t
+j
q(s) ds,
<>
(2.16-17) imply the inclusion
By (2.15) this means that z E Z. Thus under the assumptions of I the assertion is proved. IV. The case of the region V- = (a, b) x ((c, d) x (-00,0)) may be considered analogously. Referring to Assertion 1.1 completes the proof. • Recall that a sequence of points {ai: i = 1,2, ... } of a topological space X converges to a set A ~ X if every subsequence a ~ {ai: i = 1,2, ... } contains a subsequence converging to an element of the set A (this is a particular case of the notion of the convergence of sequences of sets, see §2.7 and Lemma 2.5.1). Denote by Li([a, b]) the set of all non-negative functions from LI ([a, b]). Assertion 2.2. With the notation of (2.1 - 2) let;
(2.18) AI, A2 be weakly compact subset of the set Li([a, b]); (2.19) B I , B2 be weakly compact subset of the set Li([c, d]). Then {Z( -'PI, 'P2, -'¢l, '¢2): 'PI E AI, 'P2 E A 2, '¢l E B I , '¢2 E B 2 } is a compact subset of the space Rce(U),
Proof. The proof consists in the direct reference to Assertion 2.1 and to properties of the topology of Rce(U). • Assertion 2.3. With the notation of (2.1 - 2) let conditions (2.18) and (2.19) hold;
400
CHAPTER 13
(2.20) a sequence of functions {CPij: i = 1,2, ... } ~ Lt([a, b]), j = 1,2, converge weakly to the set A j ,. (2.21) a sequence offunctions {'l/Jij: i converge weakly to the set B j .
= 1,2, ... } ~
Lt([c,d]), j
= 1,2,
Then the sequence of the spaces
converges in the space Rce(U) to the set
Proof. The proof consists in direct reference to Assertion 2.1 and to the properties of the topology of the space Rce(U). • Example 2.1. With the notation of (2.1) let CP1, CP2 E L1([a, b]). Then the set
A = {cp: cP E L 1([a,bj), CP1(t):::;; cp(t):::;; CP2(t) for almost all t E [a,b]} is weakly compact, see Chapter 5. Example 2.2. Let the assumptions of Example 2.1 hold and c > O. Then the set
A, =
{'I"
'I' E A,
i
£ }
is weakly compact. This follows from remarks of Example 2.1 and from the continuity of the linear functional
J b
f(cp) =
cp(s) ds
a
with respect to the weak topology. Example 2.3. Let a function g : [0,00) grable,
---+
1R be locally Lebesgue inte-
t
G(t) =
J
g(s) ds,
c > 0,
o
(2.22)
-00
:::;;
liminf g(t) :::;; lim sup g(t) :::;; {3 < t--+oo
t
t--+oo
t
00,
Equations and inclusions of second order. Cauchy problem theory. lim sup 2G;t)
(2.23)
t
t-+oo
~ 13 -
401
c.
By remarks of Example 2.2 the set
B = { '1"
'I' E L, ([a, b]),
I
at"
'1'( t) " fit
'1'(8) d8 " (fI-2elt'
for almost all
for all
t E [0, 1J,
t E (0,1] }
is weakly compact. For A ~ 1 and t E [0,1] put
g(~t) .
gA(t) =
Show that for every sequence Ai ~ 00 the sequence of functions 'Pi (t) = 9Ai (t) converges weakly to the set B. Start with the verification of the fulfilment of the condition
°
(2.24) for every c > there exists 8 ~ [0,1] is less than 8, then
U
> 0, such that if the measure of a set
Jl'Pi(s)1
ds
~c
u for every i Fix c that (2.25)
= 1,2, ....
> 0. By (2.22) for every a -
'TJ
'TJ
°
> there exists a number
g(t) < 13 + 'TJ t
<-
for all
t
J..I.lI
> 1 such
> J..I.1I"
Let M = sup g([O, J..I.d) and Mo = M + lal + 1131 + l. By the definition of Mo and (2.25) l'Pi(t)1 ~ Mo for all i = 1,2, ... and t E [0,1]. This implies the fulfilment of (2.24). Next, (2.24) implies the relative weak compactness of the sequence {'Pi: i = 1,2, ... }. It remains to prove that the set Bo of the limit points of this sequence lies in B. Take arbitrary 0, 'TJ E (0,1). Beginning with some i = i o, AiO > J..I.1I" By (2.25) this means that
402
CHAPTER 13
(2.26) (a - rJ)t
~
/\i
t E
~
g(\).i t ) . t
[0,1].
t
(f3 + rJ)t for all i = io, io
+ 1, ...
and
The set
B(O,,,,) = {
[0, I]}
E
is closed in the space £1([0,1]) and is convex. Therefore it is weakly closed. By (2.26) Bo ~ B(O, rJ). Since this IS true for all 0, rJ E [0,1] the set Bo lies in the set B1 = n{B(O,,,,): 0,,,, E (0, I)} = {
at
~
~
f3t for almost all t E [0, I]}.
Take an arbitrary function
J
o
~im
}ec
'->00
t
J
0
~im
.eC
'->00
J ()
t
J9(~iS)
ds
/\i
0
Ait
· \2 1 = 11m ~C
i-too
~
/\'
t.
9
U
t) dU = 1·,1m G().i \2 ~C
i-+oo
0
/\'
t.
(f3 - c)t 2 2
for every t E (0,1) (the last estimate follows from (2.23». By virtue of the arbitrariness in the choice of
+ f(x)x' + h(t, x)
=
°
In this section we point a possibility of a correct development of the Cauchy problem theory for the equation (3.1)
x"
+ f(x)x' + h(t, y)
= 0,
in the case where the function f is Denjoy integrable.
Equations and inclusions of second order. Cauchy problem theory.
403
Lemma 3.1. Let a real function f be defined and be continuously differentiable on a segment [a, b], a < b, of the real line. Then p,( {f(t) : t E [a,b], 1'(t) = O}) = O. Proof. For every c > 0 the set ME: = {f(t): t E [a, b], 11'(t)1 < c} is open by virtue of the continuity of the function 1', see Theorem 1. 7.5. By Theorem 2.6.7 the set ME: falls into the union of the family 'Y of its connected component. Let points a < (3 belong to an element U of the family 'Y. By Theorem 4.10.1 /3
If({3) - f(a)1 =
J
f'(s)ds
~ c({3 -
a).
0<
SO p,U(U)) ~ cp,(U) and p,(f(ME:» ::::; cp,(ME:) ::::; c(b - a). The inclusion {f(t): t E [a, b], 1'(t) = O} ~ ME: implies the estimate p,( {f(t) : t E [a, b], 1'(t) = O}) ::::; c(b - a), Since this is true for every c > 0 we have what was required. • Lemma 3.2. Let a function f be defined and be continuously differentiable on a segment [a, b], a < b, of the real line. Let it take values in a segment [c, d], c < d, being the domain of a Denjoy integrable real function h. Let g be a primitive of the function h, M = {t: t E [a, b]' 1'(t) = A}, 'Y be a family of connected components of the set [a, b] \ M and (3.2)
1)lgf(supr) - gf(infr)1 : r E 'Y} <
Then the function the condition (3.3)
~
00
= gf is generalized absolutely continuous and satisfies
~'(t) =
h(f(t» . f'(t)
Proof. I. If the set M is empty, we obtain what was required from Remark 4.11.1 and Assertion 4.11.1. In the general case Remark 4.11.1 implies the generalized absolute continuity of the function ~ on elements of the family 'Y. Assertion 4.11.1 implies the fulfilment of (3.3) on elements of the family 'Y. II. By Lemma 3.1 p,(f(M» = O. By the Corollary of Theorem 4.5.1 we have p,(~(M» = p,(g(f(M») = O. Enumerate the family L of all connected components of the open set U = ~ \ ~(M): L = {Ai: i = 1,2, ... }. Let 'Yi be the family of all connected component of the set [a, bj \ ~-1cI>(M) being mapped into Ai, 'Yo = ubi: i = 1,2, ... }. III. Let points 81 < 82 belong to the set M. Show that:
CHAPTER 13
404
For definiteness put 4>(sd < 4>(S2)' Notice the finiteness of the family,: of all the interval J of the family ,i, satisfying the conditions 4> (inf J) =1= 4> (sup J) (Figures 13.1 and 13.2) and J ~ [SI' S2]' (If it is not so then the function 4> has a discontinuity in points of the set [U,:] \ U,:.) Here the difference between the number of elements in the family
,t = {J:
J E ,;, 4>(inf J)
= inf( 4>( J»}
(Figure 13.2)
and the number of elements in the family
,i = {J:
J E ,;, 4> (inf J)
= sup(4)(J))}
(Figure 13.1)
is equal to 1 if the interval Ai lies between the points 4>(sd and 4>(S2)' In the opposite case the families and ,i are the same cardinality. Thus for i = 1,2, ... and Ai ~ [4>(Sd,4>(S2)] we have
,t
J
J
Figure 13.2
Figure 13.1
sup Ai - inf Ai = L {4>(sup J) - 4>(inf J): J E
+ L{4>(supJ) (3.4)
,t}
4>(infJ): J E
,i}
= L {4>(sup J) - 4>(inf J): J E
,n
= L {4>(sup J) - 4>(inf J): J E
,i,
J ~ [SI' S2]}'
Since JL(4)(M)) = 0, by (3.4)
4>(S2) - 4>(sd = JL([4>(SI), 4>(S2)] \ 4>(M» =
(3.5)
L {sup J - inf J: J
E
L, J ~ [4>(sd, 4>(S2)]}
= L {{4>(sup J) - 4> (inf J): J E ,d: i = 1,2, ... ,
Ai
=
~
[cp(Sd,4>(S2)]}
~) cp(sup J) - CP(inf J): J E
,0, J
~ [SI' S2]}'
Equations and inclusions of second order. Cauchy problem theory.
405
IV. Show that with the notation of III
Let a = {r: r E 'Y, r ~ [SI' S2]}' When we apply (3.5) to SI and S2 = sup H for H E a, we obtain:
= inf H
(3.7) L: {
a}
= L:{ O. From (3.2) there exists a number 'f/ > 0 such that as soon as 'Y' ~ 'Y and db') < 'f/, then L {
VI. Define a function
= 0 for t E
M
for t ErE T Now (3.2) implies the Lebesgue integrability of the function
~(t)
t
=
J
a
Under the assumptions of III ~(sd - ~(S2) = O. This is a direct consequence of (3.6). VIII. The function ~ is a difference of a generalized absolutely continuous function and an absolutely continuous function. Thus it is generalized absolutely continuous and VII implies that ~'(t) = 0 for almost all t E M.
CHAPTER 13
406
Therefore '(t) = cp(t) = 0 = h(f(t)) . f'(t) for almost all t E M. By I we obtain what was required. The lemma is proved. • Let real functions f and h from (3.1) be defined for all real values of arguments. Assume that: (3.8) the function
f is Denjoy integrable;
(3.9) the function h is measurable in t (for every fixed x), is continuous in x (for every fixed t), cp(t) is a non-negative Lebesgue integrable function, Ih(t, x)1 ~ cp(t) for all t, x E R In connection with the equation (3.1) consider the equation
x"
(3.10)
+ (F(x))' + h(t, x) = 0,
where the function F is a primitive of the function f. The presence of the second derivative in the equations (3.1) and (3.10) means, in particular, the continuous differentiability of every solution of each of these two equations. If the function x : [a, b] ---t IR is continuously differentiable and its derivative does not vanish on (a, b), then Assertion 4.11.1 implies that the function x solves one of this two equations if and only if it solves the another one. If here x'(a) = x'(b) = 0, then
J b
0= x'(a) - x'(b) =
x"(s)ds
a b
=
b
Jf(x(s)) . x'(s)ds + Jh(s, x(s))ds (F(x(b)) - F(x(a))) + Jh{s, x(s))ds, IF(x{b)) - F{x{a))1 ~ Jcp{s)ds. a
a
b
=
a
b
(3.11)
a
Therefore if a function x solves one of equations (3.1) or (3.10), then it satisfies the condition (3.2), because in this case the sum in the left-hand side of (3.2) does not exceed
J b
cp{s)ds.
a
Equations and inclusions of second order. Cauchy problem theory.
407
Now Lemma 3.2 implies the equivalence of the equations (3.1) and (3.10). When we put x' = y and x' + F{x) = z (that is, z = y + F{x)), we transform the equation (3.1) into the system (3.12)
{
X'
=y
y' = - f(x)x' - h(t, x),
and we transform the equation (3.10) into the system (3.13)
{
X'
= z - F(x),
Zl
= -h(t, x).
The second system locally satisfies the Caratheodory conditions. Therefore the space Z of its solutions satisfies conditions (c) and (e). The first system may be obtained from second one by the change of variables (t, x, z) -+ (t, x, z - F(x)). This change corresponds to the Lienard transformation. It keeps the fulfilment of conditions (c) and (e) So the space Z of solutions of the first system satisfies conditions (c) and (e). Now let: (3.14) functions fk' k = 1,2, ... be defined on the real line and be Denjoy integrable, Fk be a primitive of the function !k and the sequence of the functions {Fk : k = 1.2, ... } converge uniformly on every segment to a function F being a Denjoy primitive of a function f. Let Y k be a solution space of the system (3.15) and
~k
(3.16)
{
X'
= y,
y' = - fk(x)x' - h(t, x), be a solution space of the system
{
X' = z - Fk(x), z' = -h(t, x).
The sequence of the spaces ~k converges to the space Z2 of solutions of the system (3.13). This follows from classical theorems on the continuity dependence solutions on the right hand side, see §7.1. The change of variables CPk(t,X,Z) = (t,x,z - Fk(X)) transforms the space ~k into the space Yk. The sequence of the changes of variables {CPk : k = 1,2, ... } converges to the change of variables cp(t, x, z) = (t, x, z - F(x)). The sequence of inverse mappings {cp;l: k = 1,2, ... } converges to the
408
CHAPTER 13
mapping cp-l. By the Corollary of Theorem 10.5.2 the sequence 0d k) = Yk , k = 1,2, ... , converges to the space 0{Z2) = Zl of solutions of the system (3.12). We have proved: Theorem 3.1. Let conditions (3.14) and (3.9) hold. Then the sequence of solution spaces of the systems (3.15) converges to the solution space of the system (3.12). • It is naturally to understand this assertion as an assertion about the continuity of the dependence of solutions of the equation (3.1) on the function f; moreover, we consider on the set of such functions f the convergence in the sense of (3.14).
CHAPTER 14
EQUATIONS AND INCLUSIONS OF SECOND ORDER. PERIODIC SOLUTIONS, DIRICHLET PROBLEM
Here we consider a few isolated results related to the topic mentioned in the title. We do not aim to give here a full account of the topic. We try only to show that the passage to weaker (in comparison with the classical theory) restrictions on the continuity of the right hand side brings into the reasoning. Outside our consideration we leave, in particular, approachs using methods of the Algebraic Topology. Methods of the Algebraic Topology are essentially stronger than elementary arguments of Geometry and of Mathematical Analysis used here instead of them. However, our aim in this chapter (as in the book on the whole) is not to obtain maximal results at any price, but to discuss relationship of notions of topological contents that are helpful in the investigation of equations and inclusions. Our aim here is not to obtain strong final results covering all investigations in the domain in question. Our aim is to show, in particular, how to investigate discontinuous right hand sides in connection with the topic mentioned in the title. We take as a pattern the investigations from [GOl-2,HZl-2], where we meet completely classical restrictions on right hand sides of equations under consideration. The classical theory of Ordinary Differential Equations uses metric estimates as inequalities. Our topological structures allow us to pass to softer topological characteristics of corresponding relations. Results obtained in this way turn out to be more in volume for equations with continuous right hand sides too. An interesting point here is the use of asymptotic estimates of the topological structure of the solution space at infinity instead of metric estimates of the classical theory as is usual in such cases. 1. First remark on existence of periodic solutions
Consider the differential inclusion (1.1) Assume that
x" E F(t, x, x').
CHAPTER 14
410
(1.2) the right hand side F(t, x, y) = G(x, y) + H(t, x, y) is defined on the region U (with the notation of (13.1.4)) and satisfies conditions (13.1.19-25) and (13.1.6-7). As in Chapter 13 we regard the equation (1.1) as equivalent to the system
(1.3)
{
=y, y' E F(t, X, y).
X'
An approach to existence theorems for boundary values problem for an equation of the second order is developed in [GOI-2]. In this and in the next sections we compare the ideas of [GOI-2] with our theory. In order to highlight the main new ideas we restrict our discussion by only one of main results of [G02]. We do not move to a completeness of the study of all cases or to the maximal generality when this overloads the account by details being nonessential for our purposes. Theorem 1.1. Let condition (1.2) hold,
(1.4) a M
= -00,
= J~7r w(s)
b=
00, -00
~ c
~ 00, W E
Lt([O, 27fD,
ds,
(1.5) the mapping F(t, X, y) be periodic in t with the period 27f, (1.6) F(t, X, y)
~
[-w(t), w(t)] for all t E [0,27f],
(1.7) sup(U{F(t, C, 0): t E [0, 27f]})
~
0,
(1.8) inf(U{F(t, D, 0): t E [0, 27f]})
~
O.
X
E [C, D] and y E ~,
Then the inclusion (1.1) has a solution x(t), satisfying the conditions
(1.9) 7f(x) = [0,27f], (1.10) x(O)
= x(27f), x'(O)
=
x'(27f),
(1.11) C ~ x(t) ~ D and -M ~ x'(t) ~ M for all t E [0, 27f1. Proof. I. Let functions g : (c, d) x ~ -+ ~ and h : ~ x «c, d) x ~) -+ lR be continuous. With the notation of (1.2) let G(x, y) = {g(x, y)} and H(t,x,y) = {h(t,x,y)}. Then the inclusion (1.1) is the equation
(1.12)
X"
=
f(t, x, x')
Periodic solutions, Dirichlet problem.
411
with the continuous right hand side f(t,x,y) = g(x,y) + h(t,x,y). It is more convenient to pass from the equation (1.12) to the equation
x" = fl(t,X,X'),
(1.13) where
for x::::;; C, for C::::;; x::::;; D, for x ~ D.
f(t, C, y) { fl(t, x, y) = f(t, x, y) f(t, D, y)
Let a function e(t) be defined on the real line JR, be periodic with the period 27l" and be continuously differentiable. Let A > O. Theorem XII. 1. 1 of [Hal implies the existence and the uniqueness of a periodic solution z(t) = (x(t),y(t)) of the period 27l" of the system
Denote: z = l¥.x(e). By (1.7) the boundary of the region p = {(u,v): (u,v) E JR2,
X
< C, y < O}
consists of entry points (Figure 14.1). Therefore the curve z(t) either does not meet the closure of the region P, or lies inside it entirely. In the last case the function x(t) must be decreasing. This contradicts its periodicity. Thus the curve z(t) does not meet the closure of the region P. j.
l'
"
p
t
Q
,-1.
+
=r
._Lt-LL+-1-+1-1 .. +·I-J.· Ji·.. I ; 1 , 1, .,1. ! C Figure 14.1
Figure 14.2
Likewise the boundary of the region
Q = {( u, v): (u, v)
E
JR2,
X
< C, y > O}
CHAPTER 14
412
consists of exit points (Figure 14.2). Therefore the curve z(t) either does not meet the closure of Q, or lies inside it entirely. In the last case the function x(t) must be increasing. This contradicts its periodicity. Thus the curve z(t) does not meet the closure of the region Q. When we joint this two remarks we obtain that (1.14)
x(t) ~ C for all t E ~.
Likewise we obtain from (1.8) (1.15)
x(t)::;; D for all t E R
Since the function x cannot be strongly monotone the function y = x' either equals identically to zero or changes the sign. In every case y(t o) = 0 for some to E ~. Then for t E [to, to + 27r] 2'n"+to
(1.16)
ly(t)l::;;
f
211'
ly'(s)1 ds =
to
f
ly'(s)1 ds ::;; M.
0
The system
{
x'(t) = y(t), ly'(t)1 ::;; w(t)
satisfies the Davy conditions. Therefore the space
= {z: z
E
z(O) = z(27r), Im(z)
~
=
[0,27r], [C, D] x [-M, M]}
is compact with respect to the metric of the uniform convergence. Evidently it is convex. When we extend an arbitrary function (e, e') E . = a>.(e) 1[0,211']' we obtain the continuous mapping f3>. : . (see Schauder's Fixed Point Theorem 2.8.5). The function z>. solves the system (1.17)
{::::'(x- C;V) +f(t,x,y)
(see (1.14-16) in connection with the change (1.18)
11
to I),
Periodic solutions, Dirichlet problem.
413
From (1.12-16)
(1.19)
Im(z>.)
~
[C,D] x [-M,M].
The spaces Z>. of solutions of the systems (1.17) converge to the space Z of solutions of the system
(1.20) as A - t O. By (1.19) for a sequence Ai - t 0 the sequence of the solutions {Z>.i : i = 1,2, ... } converges to a solution Z of the system (1.20). From (1.18-19) (1.21 )
7I"(Z)
= [0,271"],
z(O)
=
z(271")
and
(1.22)
Im(z)
~
[C,D] x [-M,M].
Under the assumptions of I we have what was required. II. Weaken the assumptions of I and pass to the functions 9 and h satisfying the Caratheodory conditions. That is, assume the fulfilment of conditions (13.1.19-25) without their reservation in relation with y#-O (in particular, in (2.22) we assume that K ~ (e, d) x ~). Assume in addition the existence of a number N ~ 0 such that g(x, y), h(t, x, y) E [-N, N] for all t E ~,x E [C, D] and y E [-M -1, M +1]. Fix el E (e, C) and dl E (D, d). Apply the Scorza-Dragoni theorem 4.13.4 to the functions gl[cl,dl]X[-M-l,M+l] and hl[o,211"]X([Cl,dl]X[-M-l,M+lJ)' So for every i = 1,2, ... we can point closed sets P ~ [el' dd and Q ~ [0,271"], such that JL([el,d l ] \ P) < 2- i , JL([0,271"] \ Q) < 2- i (with {el,dd ~ P, {0,271"} ~ Q) and the functions gi = glpX[-M-l,M+l] and
= hIQx([Cl,dl]X[-M-l,M+lJ) are continuous. Extend the functions gi and hi to continuous functions 9i and hi in the following way according to Remark 4.13.1. Let x E [el' dl ] \ P, y E [-M - 1, M + I] and (a, (3) be a connected component of the set [el' dd \ P containing the point x. Put hi
_ gi(X,y)
x-a -a
(3-x -a
= -(3-gi({3,y) + -(3-gi(a,y).
Let·t E [0,271"j \ Q, x E [el' dlj, y E [-M - 1, M + I] and (a, (3) be the connected component of the set [0,271"j \ Q containing the point t. Put
t-a hi(t,x,y) = -(3-h i ({3,x,y) -a
(3-t
+ -(3-hi(a,x,y). -a
414
CHAPTER 14
Extend the function hi(t, x, y) by the periodicity. In order to simplify the description of the situation we have considered a concrete shape of the domain of the right hand side of the equation (inclusion). Now this complicates a little further reasonings. In order to avoid these complications and to have the possibility of a correct reference to previous material let us introduce into consideration the functions
t(X,-MJ fJi(x, y) = y) ~i(X,
gi(X, M) and
y~
if if if
t(t,X,-MJ hi(t, x, y) = ~i(t, x, y)
M,
y~M,
if if if
hi(t, x, M)
-M,
-M~y ~
y~
-M,
-M~y~M, y~M.
Evidently fJi(x,y) + hi(t,x,y) E [-2N,2N] for all t E lR, x E [cl,d 1 ], y E [-M - 1, M + 1]. By I for every i = 1,2, ... there exists a solution Z = Zi of the system (1.23)
{
=y, y' = fJi(x,y)
X'
+ hi(t,x,y)
satisfying (1.21) and Im(z)
~
[e, D]
x [-4N7r,4N7r].
By results of §§12.2 and 13.1 the solution spaces of the systems (1.23) converge to the solution space of the system (1.20) as i ~ 00. When we repeat now the final reasoning of I, we obtain the existence of a corresponding solution under the assumptions of II. By (1.6) this solution satisfies not only (1.22'), but (1.22) too. III. Omit in the assumptions of II the last requirement of the existence of the number N. For i = 1,2, ... , t E lR, x E (c, d) and y E lR put 9;(X, yJ =
and
hi(t, x, y) =
h;,
{ -;
h~t,
yJ
x, y)
if if if if if if
g(x,y)
~
-i,
- i ~ g(x, y) ~ i,
y
~
g(x,y),
h(t, x, y) ~ -i, - i ~ h(t,x,y) h(t, x, y) ~ i.
~
i,
415
Periodic solutions, Dirichlet problem. The system (1.24)
satisfies the assumptions of II. By results of §12.2 the solution spaces of the system (1.24) converge to the solution space of the system (1.20) as i ~ 00. When we now repeat the final reasoning of I, we obtain the existence of a corresponding solution under the assumptions of III. IV. Let now conditions (13.1.19-25) be fulfilled without the reservation in relation with y i= O. Fix Cl E (c, C) and d 1 E (D, d). The multi-valued functions GI[cl,d1]XIR and HI[o,21r]X([Cl,d 1]XIR) satisfy the Davy conditions. Take the approximation of §8.2. Every occurring mapping gi and hi satisfies the conditions imposed in III on the mappings 9 and h. This implies the existence of a solution Zi of the corresponding system (1.24) satisfying (1.21-22) (for Z = Zi). When we repeat now the reasoning of III, we obtain the existence of a solution Z of system (1.3) satisfying (1.21-22). V. Pass to general case. For i = 1,2, ... consider the mappings G( i
H( i
) _ {G(X,y)
x, y -
cc(G(x, 2- i ) U G(x, -2- i ))
)_{H(t,X,y) t, x, Y - cc(H(t, x, 2- i ) U H(t, x, _2- i ))
Iyl > 2- i , Iyl ~ 2- i ,
for for for for
Iyl > 2- i , Iyl ~ 2- i .
The systems (1.25)
are considered in IV. By IV there exists a solution Zi of the system (1.25) satisfying (1.21-22). The solution spaces of the systems (1.25) converge to the solution space of the system (1.3) as i ~ 00. This implies what was required. (This is quite analogous to the corresponding limit passages in III and IV.) The theorem is proved. •
2. Second remark on existence of periodic solutions In this section we consider from our point of view the main part of the proof of Theorem 2.1 of [G02]. We change metric estimates of [G02] of the behavior of the right hand side at the infinity to topological estimates of the behavior of the solution space.
CHAPTER 14
416
Assertion 2.1. Let 9 E Lt([O, 1]),
(2.1)
G(t) = J~ g(s) ds,
(2.2)
f3 > 0
and (2.3)
g(t)
~
f3t for t E [0,1].
Let a continuous function x : [p, q] -+ IR solve on the interval (p, q) the inclusion x" E [-g(x), 0], 0 < x(t) < 1 for all t E (p, q) and x(p) = x(q) = O. 7r Then q - p ~ ..JfJ. If, in addition, G(t) <
(2.4)
f3t 2
2
for t E (0,1],
7r
then q - p > -..JfJ. Proof. Since -g(x)x
~
XliX
I
0 and
q
q
j (X'(t»2 dt
= x'(t)x(t)
p
q
- j x(t)x"(t) dt
p
q
(2.5)
~
=-
p
p
q
j(X'(t))2 dt p
~
q
j x(t)x"(t) dt,
q
j g(x(t»x(t) dt
~
p
f3 j(X(t»2 dt. p
Since q
(2.6)
j(X'(t»2 dt
~
2
q
(q: p)2 j(X(t))2 dt
p
p
(see [Fi, Chapter XX, §1]), (2.5) implies the inequality 7r 2 (
q-p )
2
~
f3.
So (2.7)
q- P~
7r
V1J.
417
Periodic solutions, Dirichlet problem.
The crude estimate is obtained. Now let condition (2.4) be fulfilled. The equality in (2.7) implies the equality in (2.6). It is possible only for the function x(t) = Asin (t - p)J7j) . For this function x(t) the estimate (2.5) goes into the equality q
q
j g(x(t))x(t) dt = {3 j (X(t))2 dt. P
p
So q
(2.8)
j({3x(t) - g(x(t)))x(t) dt
= O.
p
By (2.3) the integrand is non-negative. By (2.8) g(x(t)) g(u) = (3u for all u E x([P, q]).
So G(t) =
= (3x(t), and
(3;2
for every point t E x([P, q]), which contradicts (2.4). The assertion is proved. • The change of variables ~)..(t,x,y) = (t,.Ax,.Ay),.A > 0, transforms the region x > 0 onto itself. Under this change the system (1.3) goes into the system (2.9)
{
X' =
y,
y' E .AF(t, I' f).
Denote the solution space of (2.9) by Z)... In order to estimate asymptotic properties of the inclusion (1.1) consider a sequence
such that for some (2.11)
(2.12)
kE(O,l)
Ai+l .Ai ....... ~
k
fior a 11
z. = 1 , 2 , ...
,
418
CHAPTER 14
consider the strip W =
(2.13)
~
x ({O, 1) x
~)
and (2.14) a weakly compact subset A of the set Li{[O,I]), whose elements satisfy (2.3-4) {with the notation of (2.1-2)). By Assertion 13.2.2 the set ( of solution spaces of inequalities
o ~ x" ~ -rp{x), where rp runs over the set A from (2.14), is a compact subset of Rce{W). Assertion 2.2. Let conditions (2.2), (2.10 - 14) hold. Let - 00 < Po < qo < 00, (2.15) the sequence of the spaces {{Z.dw: i = 1,2, ... } converge to the set ( as i - t 00.
Then there are numbers M > 0 and system (1.3),
C
> 0, such that if z
= (x, y) solves the
(2.16) 7r{z) = [p, q] ~ [Po, qo], x(t) > 0 for all t E (p, q), x(t o) some to E (p, q) and x(p) = x(q) = 0,
7r then q - P > ...((J
~
M for
+ c.
Proof. Assume the opposite. Then for every i = 1,2, ... we can fix a solution Zi = (Xi, Yi) of (1.3) such that: (2.17) 7r(Zi) = [Pi, qi] ~ [Po, qo], Xi(t) > 0 for all t E (Pi, qi), Xi(t i ) ~ i for some ti E (Pi, qi), Xi(Pi) = Xi(qi) = 0, qi - Pi ~ + Ci for a sequence of numbers Cl > C2
.iii
> C3 > ... , Ci
-t
O.
For every fixed i = 1,2, ... due to the convergence Aj - t 0 we can choose an index j > i such that Aj < kAi' Take the smallest of such indices j. Then
According to this reasoning select a subsequence of sequence (2.10) in order to have (with keeping ofthe notation (2.10) ) the fulfilment of the condition
Periodic solutions, Dirichlet problem.
(2.18)
k2
T
~ AHl
•
419
e0f a 11·'t = 1,2, ... < klor
instead of (2.12). Without loss of generality we can assume in addition that ti in (2.17) is the point of an absolute maximum of the function Xi (in the opposite case we change ti to a point of maximum) and that the sequence A;(!), i = 1,2, ... , monotonically increases, where
(in the opposite case we pass to a corresponding subsequence of the sequence {Xi: i = 1,2, ... }). Denote by Zi = Uh ih) the function b. Aj (;)+2 (Zj(i)). By the choice of Aj(i) we have Aj(i)Xi(ti ) > 1 and Aj(i)+lXi(ti ) ~ 1. Now (2.18) implies that:
and
Thus (2.19) 7r(Zi) = [Pi, qi] ~ [Po, qo], Xi(t) > 0 for all t E (Pi, qi), ti E (Pi, qd, k4 < Xi(t i ) ~ k < 1, Xi(Pi) = Xi(qi) = 0, qi - Pi ~ ~
+ Ci,
where
Ci
> 0 and
Ci -+
o.
The function Zi = zil(p;,q;) belongs to «ZA;)W)-+. By virtue of the compactness of the segment [Po, qo], of (2.15) and of results of §10A there are a function Z = (x, y) E U{X-+ : X E (} and a subsequence {ii : i E C} of the sequence {Zi: i = 1,2, ... } such that: (2.20) the limit to = ~~IJ}ti E [Po, qo] n 7r(z) exists,
(2.21) I ~ 7r(Zi) for ~~;ry segment I ~ (p, q) = 7r(z), beginning with some i = io E C, and the sequence {iii!: i E C, i ~ io} converges uniformly to the function Z I"
I
Condition (2.21) and the first part of (2.19) imply that:
(2.22)
(p, q)
~
(Po, qo).
CHAPTER 14
420
Condition (2.21) and the second part of (2.19) imply that: (2.23) Im(z) ~ (0, k) x ~ and x(t o) ~ k4.
Since
x EU{X-+:
X
EO, the function x solves the
o ~ (x)"
(2.24)
~
inequality
-c,o(x)
for a function c,o E A. Nw consider an arbitrary segment [tl, t 2 ] every t E (h, t2). By (2.24)
~
(p, q) such that yet) > 0 for
~((y(t))2)' + (q>(x(t)))' ~ 0 for almost all t E [tl' t2]. Here
f
u
q>(u) =
c,o(s)ds.
o
Likewise if yet)
< 0 for every t
E
(tl' t 2), then
1 2((y(t))2)'
+ (q>(x(t)))'
~ 0
for almost all t E [tl, t2]. So the maximum of the expression 1jJ(t) = t(y(t))2 + q>(x(t)) is assumed when yet) = O. Then 1jJ(t) = q>(x(t)). Since the function q>( u) is non-decreasing and the point to is a point of absolute maximum of x, the maximum of 1jJ(t) is assumed at t = to. Therefore the curve z(t) does not leave the set
P = {(u, v): 0 < u ~ k,
v2
2 + q>(u)
~
q>(x(to))} ~
w.
The set PI = Pu {CO, v): v 2 ~ 2q>(x(t o))} is compact. From the closed ness of the set Gr(z) ~ [Po, qo] X PI (see §10.3) limt--+p x(t) = limt--+q x(t) = o. By (2.21)
By Assertion 2.1 this is impossible. The contradiction obtained gives what was required. The assertion is proved. • Assertion 2.3. Let conditions (2.10 - 15) hold, (2.25) a > 0, a- t
+ {3-t = 2,
421
Periodic solutions, Dirichlet problem.
(2.26) a sequence of numbers {/Li: i
Pi: i = 1,2, ... }), (2.27)
~
WI =
= 1,2, ... }
x ((-1,0) x
satisfy (2.10 - 12) (as
~),
(2.28) the sequence {(ZlLi )Wl : i = 1,2, ... } converge to the solution space of the system {
as
2
~
X'
=y,
y' E [0, -ax]
00.
Then there is a number C > 0 such that every solution z = (x, y) of (1.3) with
(2.29) 7r(z)
= [0,27rJ,
x(O)
= x(27r),
y(O)
= y(27r)
satisfies the condition:
(2.30) either x(t)
> -c
for all t E [0,27rJ, or x(t)
for all t E [0, 27r].
Proof. Without the loss of generality we can assume that the function F(t, x, y) is periodic in t with period 27r. Put Po = -27r and qo = 47r. Apply Assertion 2.2 and fix corresponding values of M and c. Now fix 'T/ > 0 such that 7r 7r c --===>--Ja+'T/ ..;a 2' After the change of variables x ~ (- x) apply Assertion 2.2 with f3 = a + 'T/. Fix corresponding numbers MI and CI' Put C = M + MI' Assume now that condition (2.30) is not fulfilled for a solution z = (x, y) of the problem (1.1), (2.29). We can assume that the solution is extended by the periodicity. Fix intervals (p, q) and (PI, qI) such that x(t) > 0 for t E (p, q), x(p) = x(q) = 0, x(t) ~ M
for some point
<0
t E (p, q),
for t E (PI, qd, x(pd = x(qd = 0, x(t) ~ -MI for some point t E (PI, qd. x(t)
Then
422
CHAPTER 14
Therefore
•
This is impossible. The assertion is proved. With the notation of (2.1012), (2.14), (2.26) let
<>
y=h(x)
°°
(2.31) I < and real functions a(x) < < b(x) be defined and be continuous on the real line and Ix E (a(x), b(x)) for every x E 1R,
y=a(x) Figure 14.3
(2.32) the sequence of functions
converge weakly on the segment [0, I] to the set A, (2.33) the sequence of functions
converge weakly on the segment [0, I] to the function cp == 0, (2.34) the sequence of functions
converge weakly on the segment [-1,0] to the function cp == 0, (2.35) the sequence of functions
converge weakly on the segment [-1,0] to the set
{cp: cp
E L1([-I, 0]),
°
~
cp(t)
~
-at}.
For the convenience of the comparison with previous reasonings put
Periodic solutions, Dirichlet problem.
(2.36)
423
F{t, x, y) = [a{x), b{x)].
Assertion 2.4. Let conditions (2.10 - 14), (2.25 - 26), (2.31 - 35) hold and r > 0. Then with the notation of (2.36) there exists a number C 1 > such that for every solution z = (x,y) of the problem (1.3), (2.29) with min{lx(t)l: t E [0,271"]} < r the estimates Ix(t)1 ~ C 1 and ly{t)1 ~ C 1 are true for all t E [0,271"]. Proof. As in the proof of Assertion 2.2 we pass to sequences {Ai : i = 1,2, ... } and {/Li: i = 1,2, ... } satisfying (2.18). I. By Assertion 13;2.3 we have the fulfilment of (2.15) with ( = {Z(cp,O,O,O): cp E A} and of (2.28). Thus we can use Assertion 2.3. Fix a number C > according to Assertion 2.3. II. Prove that there exists a number Co > such that for every solution z = (x, y) of the problem (2.1), (2.29) with min{lx(t)l: t E [0,271"]} < r the estimate ly(t)1 ~ Co is true for all t E [0,271"]. Assume the opposite. We can construct a sequence of solutions Zi = (Xi, Yi), i = 1,2, ... , of the problem (2.1), (2.29) such that:
°
°
(2.37)
°
min{lxi(t)l: t E [0,271"]} < r for every i = 1,2 ...
and limi-+oo max{IYi(t)l: t E [0,271"]} = 00. If it is necessary we pass to a subsequence and we obtain the fulfilment either of the condition
Xi{t) > -C for all i = 1,2 ... and t E [0,271"] or of the condition
Xi(t) < C for all i = 1,2 ... and t E [0,271"]. Assume for definiteness the fulfilment of the first of this two conditions. The second case may be considered in an analogous way. III. Show that for every number Ll > there exists a number M > such that for every i = 1,2 ... and t E [0,271"], if Xi(t) < L 1 , then IYi{t)1 ~ M. Let Mo = max{la(x)l, b{x): x E [-C, L 1 ]} and M = 271"Mo. If the set {t: t E JR, Xi{t) < Ld is nonempty then it falls into the union of the family of pairwise disjoint intervals. Let (a, f3) is one of such intervals and p E (a, f3) is a point of the minimum of the function Xi on (a, f3). Then Yi (p) = x~ (p) = and for every point t E (a, f3)
°
°
°
t
IYi(t)1
~
J
Mods
p
~
M,
424
CHAPTER 14
which was required. IV. Show that limi-+oo mi = 00, where with the notation of II mi = max{xi(t): t E [0, 27r]}. If it is not so then there exists a number L > 0 such that the set A = {i: i = 1,2, ... , mi < L} is infinite. By III the infiniteness of the set A contradicts our choice of the sequence {Zi: i = 1,2, ... ,} in II. V. Let ti be a point of the maximum of the function Xi' By IV limi-+oo Xi(t i ) = 00. Without the loss of generality we can assume in addition that Xi(t i ) > >'1 1 for all i = 1,2, ... (if this is not fulfilled for the initial sequence we pass to a corresponding subsequence). VI. For i = 1,2, ... put j(i) = sup{ m: m = 1,2, ... , >.:;/ < Xi(t i )}. By V limi-+oo j(i) = 00. By the definition of j(i) we have >'j(!)+1 ~ Xi(ti). Denote by Zi = (Xi, Yi) the function .6.,xj(il+2 (Zi)' Then (as in an analogous situation in the proof of Assertion 2.2) k4
>'j(i)+2 < Xi ti = Aj(i)+2Xi ti < -,-A
( ) '
( )
Aj(i)
~
>'j(i)+2 -,-Aj(i)+1
~
k
.
VII. Show that for every c > 0 there exists an index io such that Yi(t) < c for all t E ~ and i = i o, io + 1, .... Assume the opposite. Let Pi < qi be the endpoints of a connected component of the open set I = {t: t E ~, Xi(t) > 0, Yi(t) > H (the situation when I = ~ contradicts the periodicity of the function Xi), (3i(X) = >'j(i)+2 b (>.j(!)+2 X), By the last assumption in II and by III the set {Yi(t): i = 1,2, ... , Xi(t) = O} is bounded. So IYi(Pi)1 ~ ~c, beginning with some i = 1,2, ... . For t E (Pi, q;] t
Yi(t) = Yi(Pi)
+ / y~(s)ds ~ ~ + / Pi
=
3: + /
t
{3i(Xi(S))ds
~
=
4" +
t
3: + ~ / (3i(Xi(S))X~(s)ds 4" + J Pi
J
:i:i(t)
2 ~
>.j(i)+2b(Xi(S))ds
Pi
Pi
3c
t
:i:i(p;}
2 ...
{3i(u)du =
3c
2 ~
{3i(u)du.
0
Periodic solutions, Dirichlet problem. By (2.32) the last integral tends to zero as i -
00.
425
So
2/ f3i(u)du < 4'c 271"
~
o
3:
beginning with some i = i o. Therefore fh(t) < + ~ = c for all t, which was claimed. VIII. The estimate (2.37) and the assumption in II concerning C imply the equality limi-+oo min{xi(t): t E [0,27r]} = O. Since 271"
SUp{Xi(t): t E [0,27r]} ::;; min{xi(t): t E [0,27r]}
+/
yt(s)ds
o
(see notation in §4.2), VII imply that SUp{Xi(t): t E [0,27r]} - O. This contradicts the estimate SUp{Xi(t): t E [0,27r]} > k4 > 0, which follows from VI. The assumption in II has led to a contradiction. Hence the aim, pointed out in II, is achieved. Putting now C 1 = 27rCo + r, we obtain what was required in the full volume by virtue of the estimate t2
IXi(t2) - Xi(t 1 )1::;; / Yi(s)ds tJ
The assertion is proved. • Theorem 2.1. Let conditions (2.10 -14), (2.25 - 27), (2.31- 36) hold. Let: (2.38) g : IR X 1R2 continuous function;
IR be a
(2.39) g(t,x,y) E F(t,x,y) for all t, x, y E IR; (2.40)
(2.41) E IR;
-00
< C < D < 00;
c
D
get, C, 0) ~ 0 for all
t
(2.42) get, D, 0) ::;; 0 for all E IR (see Figure 14.4 in connection with conditions (2.4042)).
t
Figure 14.4
CHAPTER 14
426
Then the problem (2.29) for the equation x" = g(t, x, x')
(2.43)
has a solution satisfying the condition (2.44) x(t) E [C, D] for some t E [0,271"]. Proof. I. By Assertion 13.2.3 we have the fulfilment of (2.15) with ( = {Z(rp,O,O,O): rp E A} and of (2.28). Thus we can use Assertions 2.3-4. II. Fix C 1 > according to Assertion 2.4 for r = ICI + IDI. Put C 2 = C 1 + ICI + IDI + 1. Evidently there exists a number II E (l,0) such that
°
( C+D) 2 < b(x)
a(x) < 11 x for all x E R III. Put
_ {max{min{b(X)'9(t' -C2 , y)}, a(x)} f(t, x, y) g(t, x, y) max{min{b(x), g(t, C2 , y)}, a(x)}
for x ~ -C2 , for - C2 ~ X for x ~ C2 •
~
C2 ,
The continuity of the functions a(x) and b(x) in (2.31), (2.36) and (2.39) imply the existence of a number M > (1 - ll)C2 such that
If(t,x,y)1 < M
(2.45)
for all
t,x,y E R
For>. E [0,1) consider the system:
(2.46)
{::
:~; _ A)l, (x _ c; D) + V(t, x,y).
Notice that
( C+D) 2 +>'f(t,x,y)
(2.47) (1->.)1 1 x -
E
F(t,x,y) for all
t,x,y
Assume that
(2.48) a function ~(t) = (p(t), q(t)) is continuous, p(O) = p(271"), q(O) = q(271").
7I"(~)
= [0, 271"j,
E R
427
Periodic solutions, Dirichlet problem.
By Theorem XII.l.1 in [Ha] there exists a number N>. > 1 such that for every function ~ satisfying (2.48) there exists an unique solution 'l/JI (0 of the system
(2.49)
{:: :
~; _ >')1, (x _c; D) + >.J(t,p(t), q(t)),
on the segment [0,27T] with coinciding values in the endpoints, moreover
II'l/JI(OII
~ N>.M.
Denote by 4> the set of solutions of the system
on the segment [0,27T] bounded in the norm by the number N>.M and with coinciding values in the endpoints. 4> is a convex compact. We have defined a continuous mapping 'l/JI : 4> --+ 4>. Now define the second continuous mapping 'l/J2 : 4> --+ 4>, 'l/J2(P, q) = (PI, q), where
P(t) - s(p) + C { PI(t) = p(t) p(t) - i(p) + D
s(p) ~ C, s(p) > C i(p) ~ D
if if if
and
i(p) < D,
with i(p) = infp([O, 27TJ) and s(p) = supp([O, 27TJ). Apply Schauder fixed point theorem 2.8.5 to the mapping 'l/J2'l/JI : 4> We obtain the existence of a solution ~ = (p, q) of the system
(2.50)
{:: Xl
:~;
- >.)/,
=
+ a,
X
satisfying the condition x(O)
(x _
c; D) +
= X(27T),
y(O)
(2.51) a > 0 and then s(x) < C, (2.52) a = 0 and then s(x) ~ C and i(x) (2.53) a < 0 and then i(x) > D.
>.J(t,
= y(27T),
x"
--+
4>.
V),
where
~ D,
Consider the case (2.51). Let to E [0,27T] be a point of the maximum of the function x(t). Then y(t o) = 0 and a = C - x(t o). By the second equation of (2.50)
(
x"(tO) = y'(to) = (1 - A)ll x(to) -
C+D) 2 + A/(to, C, 0) > o.
428
CHAPTER 14
This contradicts the choice of the point to. Thus the case (2.51) is impossible. Likewise the case (2.53) is impossible too. Thus condition (2.52) holds. We have the solution (x, y) of the problem (2.29), (2.44) for the system
Now (2.54) and the choice of the Cl and C2 imply that
Solution spaces of the systems (2.54) converge to the solution space of the system (2.56)
{
X'
= y,
y' = f(t, x, y)
as >. ~ 1. Previous remarks imply the existence of a solution of the problem (2.56), (2.29), (2.44). By our definition of f the pointed solution solves the • initial equation too. The theorem is proved. Compare the obtained result with results of [G02j and show at once one of possibilities of its application to the investigation of particular equations. Assume that the functions a and bin (2.31) satisfy the condition lim sup a(x) :1:-+-00
(2.57)
~ 0,
lim sup b(x)
X
:1:-+00
liminfa(x) ;;;:: -a, :1:-+00 x . 2A(x) and either hm sup - - 2x--+oo x
liminf b(x) ;;;:: -(3, :1:-+-00
or
X
2B(x) lim sup - - 2 - < (3,
A(x) = -
X
:1:-+-00
J
J
o
0
:I:
where
~ 0,
X
a(s) ds, B(x) = -
:I:
b(s) ds
(we follow [GOl-2], when we introduce these conditions). Put: (2.58) bl(x) = b(x) + bo(x) and al(x) = a(x) - bo(x), where bo(x) = clxl(sin2 (lnlxl»I:l:I, c> 0 and F(t,x,y) = [al(x),b 1 (x)].
Periodic solutions, Dirichlet problem.
429
For this mapping F the system (2.9) with A = (27fk)-1, k = 1,2, ... , takes the form:
As clxl(sin2(ln Ixl»27rklxl --t 0 monotonically almost everywhere on the segments [-1, 0] and [0, 1] as k --t 00. Therefore they converge weakly to the function (J == O. Remarks of Example 13.2.3 and Theorem 2.1 imply: Assertion 2.5. Let conditions (2.38), (2.40 - 42), (2.57- 58), (2.32 - 35) hold, al(x) ~ g(t,x,y) ~ b1 (x) for all t,x,y E R. Then the problem (2.29), (2.43 - 44) has a solution. • The following assertion is a slightly weakened version of the main result of [G02]. Theorem 2.2. Assume that a function 9 : IR --t R is continuous, inf g(R) = -00, sup g(R) = +00, conditions (2.2) and (2.10) hold and
(2.59)
. g(x) hmsup-x-+-co x
~
lim sup g(x) x-+co x
a,
~ {J,
(2.60) for the function x
G(x)
=
J
g(s)ds
o
at least one of the two estimates
. 2G(x) 1Imsup--2x-+-co x
or
lim sup x-+co
2B~x) < {J X
is true. Then for every continuous function h : [0,27f]
(2.61)
x"
+ g(x) + h(t)
--t
R the equation
= 0
has a solution satisfying conditions (2.29), (2.44). Proof. The proof uses the general scheme of reasoning of [Go2], but it does not repeat its technically complicated parts. The boundedness of the function h and conditions imposed on the function g imply the existence of two points A, B E R such that g(A) ~ -h(t) ~ g(B) for all t E R I. If A = B here, then h(t) == -g(A) and the function x(t) == A is a required solution of the equation (2.59).
430
CHAPTER 14
°
°
II. Let A > B. Since g(A) + h(t) ~ and g(B) + h(t) ~ for all E [0,27rj we obtain what was required from Theorem 1.1. III. Let A < B. If the function 9 is unbounded from below on [B, 00 ), then there exists a point C > B such that g(C) ~ -h(t) ~ g(B) for all t E [0,27rj and we obtain what was required from II. Next, if the function 9 is unbounded from above on (-00, Aj, then there exists a point C < A such that the inequality g(A) ~ -h(t) ~ g(C) holds for all t E [0,27r] and one time more we obtain what was required from II. lt remains to prove our theorem under the additional assumption t
(2.62)
limsupg(x)
< 00
and
lim inf g(x) :1:-+00
x--+-oo
>
-00.
Let
+ Iinf g([B, 00))1 -00, ADI + sup{lg(x)l:
r = sup{lh(t)l: t E [0,27r]}
+ Isupg« 1=
A ~ x ~ B}
+ 1,
- min { ~, ~ } ,
-r
for x ~ A, for A ~ x ~ B, for x ~ B,
+r
for x ~ A, for A ~ x ~ B, for x ~ B.
-r { a(x) = min{ -g(B), 0, lB} . 1~~
min{ -g(x), 0, lB} - r max{ -g(x), 0, lA} + r { b(x) = m:x{ -g(A), 0, lA} . ~=~
Let gdt, x, y) = {-g(x) - h(t)}. For these definitions (and for gl as g) conditions (2.31), (2.36) and (2.39) hold. The fulfilment of the first two conditions in (2.57) follows from the equalities
1I· ma(x) - - = 1·I m-r -=
x-+-oo
and
X
:1:-+-00
X
°
. b(x) . r hm - = hm - =0. :1:-+00
X
%-+00
X
The fulfilment of the third and of the fourth conditions in (2.57) follows from (2.59). The fulfilment of the last condition in (2.57) follows from (2.60). From the inequality a(x) ~ -g(x) - h(t) ~ b(x) our assertion is a • consequence of Theorem 2.1.
431
Periodic solutions, Dirichlet problem.
The addition of the term bo in Assertion 2.5 (see (2.58)) in the majorant leads from conditions of [Go2] (see (2.57)), namely: Example 2.1. Let hypotheses of Theorem 2.2 hold and c E R Then the equation x" + g(x) + cx(sin2 (ln Ix!))IXI + h(t) =
°
has a solution satisfying (2.29) and (2.44). This is a direct consequence of Assertion 2.5. In Theorem 2.1 we considered an equation with continuous right hand side. This is made in order to attract attention to the used limit passages in the space Rce(U). Now give a maximal statement: Theorem 2.3. Let conditions (2.31), (2.10 - 14), (2.2), (2.32 - 35), (1.2), (2.40) hold, F(t,x,y) ~ [a(x),b(x)] for all t,x,y E~, inf{U{F(t, C, 0): t E [0, 27r]}} ~ 0, sup{U{F(t,D,O): t E [0,27r]}} ~ 0, Then the problem (2.29) for the inclusion x" E F(t,x,y) has a solution satisfying (2.44). Proof. We refer to Theorem 2.1 and next we repeat the arguments of the steps II-V of the proof of Theorem 1.1. • 3. Upper and lower solutions of the inclusion x" E F(t, x)
Our aim in this section is to show how we can extend the method of upper and lower solutions to a differential inclusion (3.1)
x" E F(t, x).
Here we assume that: (3.2) values ofthe mapping Fare (nonempty) segments or one point subsets of the real line ~; (3.3) for every fixed t E ~ the mapping Ft(x) = F(t, x) is upper semicontinuous; (3.4) the mapping FX(t) = F(t,x) is measurable for every fixed x E ~. Denote by W the domain of the function F. By Theorem 5.12.3 and (3.4):
CHAPTER 14
432
(3.4') there exists a single valued function f : W ---+ IR such that f(w) E F(w) for all w E Wand the function !,,(t) = f(t, x) is measurable for every fixed x E R As before we regard the inclusion (3.1) equivalent to the system x'-y y' ~ F'(t,x).
{
(3.5)
Consider the boundary values problem for the inclusion (3.1) with
x(o) = a, x(7I") = b.
(3.6)
Following [HZl-2] in the generalization of the notion of a lower solution assume that: (3.7) 0'1, ... , a r is a family of functions defined on the segment [0,71"] and having absolutely continuous derivatives, a(t) = max{ 0'1 (t), .. . ,ar(t)} and for almost every point t E (0,71") there exists a number i = 1, ... , r such that a(t) = ai(t) and a~'(t) ~ inf F(t, a(t)). Assume also that:
0'(0)
(3.8)
~
a
and
0'(71")
~
b.
Respectively, assume that: (3.9) {31, ... , {3. is a family of functions defined on the segment [0,71"] and having absolutely continuous derivatives, {3(t) = min{{31(t), ... ,{3.(t)} and for almost every point t E (0,71") there exists a number i = 1, ... , s such that {3(t) = {3i(t), {3?(t) ~ supF(t, {3(t)),
{3(0)
(3.10)
Denote by that
~ a
and
{3(71")
~
b.
A the set of all measurable functions h : (0,71")
f
---+
[0,00) such
11"
t(7I" - t)h(t) dt < 00.
o
Assume that: (3.11) there exists a function h E all (t, x) E W.
A such
that F(t, x) ~ [-h{t), h{t)] for
Theorem 3.1. Let conditions (3.2 - 4) and (3.7 - 11) hold,
Periodic solutions, Dirichlet problem.
433
(3.12) a(t):::;; (3(t) for all t E [0,7r], (3.13) the set WI = {(t,u): t E (0,7r), a(t):::;; u:::;; (3(t)} lie in W.
Then the problem (3.1), (3.6) has a solution Xo satisfying the condition (3.14)
a(t) :::;; xo(t) :::;; (3(t)
Ix~(t)1
(3.15)
for all
b-a :::;; 7r
t E (0,7r),
+ 'Y(t)
for almost all t E (0, 7r), where t
'Y(t) =
~ j sh(s) o
...
ds
+ ~ j(7r -
s)h(s) ds.
t
Proof. 1. Consider the case of a single valued function F. That is, with the notation of (3.4') F(t,x) = {J(t,x)} for every point (t,x) E W. Pass to the function
g(t, x) = {
f(t, a(t)) + arctan(x - a(t)) f(t, x) f(t, (3(t)) + arctan(x - (3(t))
if if if
x:::;; a(t), a(t) :::;; x :::;; (3(t), x ~ (3(t)
defined in the strip (0,7r) x JR. II. Show that with the notation of I every solution Xo of the problem (3.6) for x" = g(t, x) satisfies (3.14). Prove the fulfilment of the inequality xo(t) ~ a(t). Assume the opposite. Let tl be a point of the maximum of the function a(t) - xo(t). By (3.8) tl i= 0, 7r. By the last assumption a(td - xo(td > 0. Let Mi denote the set of all points t E (0,7r) such that ai(t) = a(t). For s E [0,7r] put l(s) = {i: i = 1, ... ,r, s E [Mi n (-oo,s)]} and r(s) = {i: i = 1, ... , r, s E [Mi n (s, oo)]}. Evidently these sets are nonempty. At every point s E (0,7r] the function a has the left derivative p(s) = inf{a~(s): i E l(s)} = inf{a~(s): i = 1, ... ,r, SEMi}. At every point s E [0, 7r) the function a has the right derivative q( s) = sup{ a~ (s) : i E r(s)} = sup{a~(s): i = 1, ... ,r, s EM;}. For s E (0,7r) we have p(s) :::;; q(s). In addition, if l(s) n r(s) i= 0 then p(s) = q(s). If a point sEMi is not the right endpoint of a connected component of the complement to M i , then a~ (8) = p( 8). If a point 8 E Mi is not the left endpoint of a connected component of the complement to M i , then aas) = q(8). This implies that the equality p(8) = q(8) = a'(8) holds on
CHAPTER 14
434
the segment [0,71"1 everywhere, except an at most countable set of values of the argument s. Fix c > 0 and m > 0 such that (t i - C, tl + c) ~ (0,71") and a(t) - xo(t) > m for t E (ti - C, tl + c). Show that: A. The function h(t) = q(t) - x~(t) - t . arctan m is non-decreasing on the interval (t i - C, tl + c). First prove that: B. If an interval (e, d) ~ (t i - C, tl + c) is covered by a family {Mi i = 1, ... , rl} then the function h is non-decreasing on (e, d). For (c, d) = (tl - C, tl + c) and rl = r B implies A. The validity of B will be proved by an induction on rl. C. Ifrl = 1, the interval (e,d) lies in MI. Then a'l(c,d) = a~l(c,d) = ql(c,d) and h'(t) = a~'(t) - f(t, a(t» -arctan(xo(t) -a(t» -arctan m ~ 0 for almost all t E (e, d), what was required. D. For rl = 2, ... ,r consider an arbitrary connected component (CI' dd of the set (e,d) \ M r1 . Since the interval (el,dd lies in the union U{Mi: i = 1, ... , rl - I}, by the inductive hypothesis the function h is nondecreasing on (el' dd. Here dl E Mrl and a~l (dd
- x~(dd - d l arctan m ~ p(dd - x~(dd - d l arctan m = lim sup (q(s) - x~(s) - s . arctan m) S ..... dl,S
~
q(cd - x~(cd - CI . arctanm ~ a~l (cd - x~(ed - el arctan m. When we repeat the argument of C (with the change of al to arJ, we see that the derivative of the function a~l (t) - x~(t) - t . arctan m on the set Mrl n (e, d) exists almost everywhere and is non-negative. Now the estimate
(for every component (el' dd of the set (c, d) \ M r1 ), the absolute continuity of the function a~l (t)-x~(t)-t·arctan m, and Assertion 4.7.1 imply that the function a~l (t) - x~ (t) - t· arctan m is non-decreasing on the set (e, d) n Mrl . To complete the reasoning notice that limsup q(s) :::;; p(sd :::;; a~ (sd :::;; q(sd = liminf q(s) 8-+81,8<81
SI E M r1 · From A the function q(t) (tl - £, tl + c).
1
8-+81,8>81
for
x~(t)
is increasing on the interval
435
Periodic solutions, Dirichlet problem.
Since tl is a point of the maximum of the function a(t) - xo(t),
q(td - x~(td ~ p(td - x~(td ~
o.
So q(t) - x~(t) > 0 for t E (tl' tl + c). This means that the function a( t) - Xo (t) is increasing on the interval (tl' tl + c). Therefore the point tl cannot be a point of the maximum of this function. The contradiction • obtained gives what was required. Likewise xo(t) ~ (3(t). III. With the notation of I associate to an arbitrary continuous function u : [O,7rJ - IR. the function
t
Tu(t) =a - -(b - a) 7r
(3.16)
t
~
7r-tj sg(s,u(s)) + -7ro
ds
+:;t j (7r -
s)g(s,u(s)) ds.
t
(Notice that the function g(s, u(s)) is measurable. By (3.11) the integrands are Lebesgue integrable. The continuity of the dependence of the integral on limits of integration implies the continuity of the function Tu). Let Ui - u. We have:
ITui(t) - Tu(t)1 =
~
t
j(7r - t)s(g(S,Ui(S)) - g(s,u(s))) ds o ~
+j
(7r - s)t(g(s, Ui(S)) - g(s, u(s))) ds
t
t
~ ~ j(7r -
t)slg(S,Ui(S)) - g(s,u(s))1 ds
o
+~j
~
(7r - s)tlg(s, Ui(S)) - g(s, u(s))1 ds
t t
~ ~ j(7r -
s)slg(S,Ui(S)) - g(s,u(s))1 ds
o
+~
J ~
(7r - s)slg(s, Ui(S)) - g(s, u(s))1 ds
t
~~
J ~
(7r - s)slg(s, Ui(S)) - g(s, u(s))1 ds.
o
CHAPTER 14
436
Since the argument t is absent in the last expression and the expression itself tends to zero as i ---t 00, IiTUi - Tuli ---t 0. Thus the mapping T is continuous with respect to the norm of the uniform convergence. The norm of the image of every function is bounded from above by the number 7r
lal + Ibl + ~ j(-Tr -
s)sh(s) ds
+ 71"3 < 00.
o
Differentiation of (3.16) in t gives
b- a + ,(t) + 71"2 Idtd Tu(t) I ~ -71"-
(3.17)
for every continuous function u for almost all t E [0,71"] where the function ,(t) is defined in (3.15). If a(t) ~ u(t) ~ (J(t) for all t E [0,71"] then b- a + ,(t). Idtd Tu(t) I ~ -71"-
The function ,( t) is integrable. So the image of the mapping T is relatively compact. By Schauder's fixed point theorem 2.8.5 it has a fixed point Xo: t
t - a) xo(t) = a - -(b 71"
7r
7I"-tj +- sg(s,xo(s)) ds 71" o
+ -t j (71" - s)g(s,xo(s)) ds. 71" t
Then the function Xo satisfies (3.1), (3.6). The fulfilment of (3.15) follows from the second part of (3.17) and from the definition of 9 in 1. IV. By III and II under the additional assumptions of I the theorem is proved. V. Pass to the general case. Theorem 4.12.3 implies the existence of measurable functions A(t) and B(t) satisfying the conditions: A(t) E F(t, a(t)), For i
B(t) E F(t, (J(t)).
= 1, 2, . .. define the function gi (t, x) in the following way. Let
t E (0,71"). Let (p, q) be a connected component of the set
IR \ ({k2- i
:
k
= O,±1 ± 2, ... } U {a(t),{J(t)}).
437
Periodic solutions, Dirichlet problem.
For u E [p, qj put
q - u A(t) q-p q-u -A(t) q-p
+u-
P B(t) q-p u-p + -q-p J ( t , q)
q - u J(t,p) q-p
+u-
P A(t)
q-p
q - u J(t,p) + u - P B(t) q-p q-p q-u u-p -B(t) + - J ( t , q) q-p q-p q-u u-p -J(t,p) + - J ( t , q) q-p q-p
if
p = aCt), q = (J(t),
if
p = aCt), q i= (J(t),
if
q = aCt),
if
p i= aCt), q = (J(t),
if
p = (J(t),
in other cases
(with the notation of (3.4)). Apply the result ofIV to the problem (3.18)
x"
= 9i(t, x),
x(O)
= a,
x(11")
= b.
Let Xi be a corresponding solution of the problem (3.18). Let Zi denote the solution space of the system X' =
{
y'
y,
= 9i(t,X)
The sequence {Zi: i = 1,2, ... } converges to the solution space of the system (3.5) as i --+ 00. By (3.14-15) on an arbitrary segment [e, dj ~ (0,11") the graph of the function (Xi, xD lies in the compactum
{ (t, u, v), c'; t.; d, a(t) .; u .; (3(t),
Ivl : :;
b- a
-11"-
1
+ :;;:
J d
o
shes) ds
+
J~
(11" - s)h(s) ds
}
.
c
Therefore the sequence {(Xi, x~): i = 1,2, ... } contains a subsequence {(xi,xD: i E A} converging uniformly on every segment I ~ (0,11") to a solutions (xo, Yo) of the system (3.5). By (3.15) and by the Lebesgue integrability of 'Y the sequence {Xi: i E A} is equicontinuous. Therefore the sequence {Xi: i E A} converges uniformly on the segment [0,11"]. So we can extend the function Xo to a continuous function on the entire segment [0,11"] by putting xo{O) = a, xo(11") = b. The theorem is proved. •
438
CHAPTER 14
Theorem 3.2. Let conditions (3.2 - 4) hold with the mapping F defined for i E (0,71") and x > 0. Assume that:
(3.19) k > 1 and for every compactum K ~ (0,71") there exists a number E> such that F(i,x) ~ (-00, -Px] for all i E K and x E (O,E];
°
(3.20) , < 1, M > 0, hi E i E (0,71") and x ~ M;
A and
F(i, x) ~ [_,2X - hi (i), (0) for of all
(3.21) for every compactum K ~ (0, (0) there exists a function h2 E such that F(i, x) ~ [-h2(i), h 2(i)] for all i E (0,71") and x E K.
A
Then problem (3.1), x(O) = 0, x(7I") = 0, has a solution: Proof. I. Fix a number k > 1 according to (3.19). II. Condition (3.19) implies the existence of a twice differentiable function al : [0,71"] --t [0,(0) such that:
supF(i,x):::;; -Px for all i E (0,71") and x E (O,al(i)]; a~(i)
°
> for all i
E
[0, V U e;, 71"].
Fix an arbitrary kl E (l,min{2,k}). Select a number A2 E (0,1] such that
where Ct2(i) = A2cos(kl(i - ~)). For this choice supF(i,x):::;; -Px for all x E (0, Ct2(i)] and i E ((1 - 11 H, (1 + 11 H)· For a rather small Al E (0,1] we can select points i l E (0, t) and i2 E (2;, 71") such that: Ctl (i) ~ Ct2(i) for all i E [0, U [i 2 , 71"]
iIl
and
(Xl(
t)
--~-.--.--.-
Ctl(i) :::;; Ct2(i) for all
i E [ii, t 2 j,
o
-------_ ... _---_.. - - _._..._-_.- ...
Figure 14.5
------.:~::::.::.-=-
439
Periodic solutions, Dirichlet problem.
For this choice Ct~(t)
- inf F(t, Ctdt)) > 0 for all t E [0, tIl u [t 2 , 7I"j,
Ct~(t)
- infF(t,Ct2(t)) > 0 for all t E [t l ,t2j.
That is, (3.7) holds. Next, Ct(O) = CtI(O) = 0, Ct(7I") = CtI(7I") = 0 for Ct = max{CtI,Ct2}. That is, (3.8) holds with a = 0 and b = O. III. Let 'flo = sup Ct«O, 71")). Fix M > 'flo, , and hI according to (3.20). Fix h2 according to (3.21) for K = ['flo, Mj. Let h = hI + h2· For this choice F(t, x) ~ (-,x 2 - h(t),oo) for all t E (0,71") and x E ['flo, 00). The function
f t
sin(ft)
r(t) = -
,
(3.22)
f
f t
cos(ft) h(s) cos(fs) ds + ,
0
.
h(s) sm(fs) ds
0
t
=
h(s) sin(ft - ,s) ds
o
'
satisfies the equation r"(t) + ,2r(t) = -h(t) for almost all t E (0,71"). The integrand expression in the last term of (3.22) is non-negative. Therefore the number A = r(7I") is non-negative too. Let 71" 71" 71"(1-,) cp = - - -, = 2 2 2 and
B=~. smcp
The function rl(t) = B sin(ft+cp) solves the equation x" +,2X = O. For t E (0,71") the point,t+cpbelongs to (cp,7I"-CP). Thereforerl(t) > 'flo = rl(O). Thus thf' function Xo = r + rl satisfies the conditions x~(t)
- supF(t, xo(t))
~ x~(t)
+ ,xo(t) + h(t)
= 0
and
xo(t)
~ 'flo
for all t E [0,7I"j. IV. Construct a sequence of functions {Xi: i = 0,1,2, ... }. The function Xo is already constructed. For the inductive construction it will be important that (3.23)
X~/_l(t) ~
supF(t,Xi_l(t))
for almost all
t E (0,71"),
CHAPTER 14
440 X~/_l (t) ~
(3.24)
a(t)
for all
t E (0,7r),
(3.25) with TJi = 2- iTJo. Theorem 3.1 implies the existence of a solution Xi of (3.1) such that
Xi(t)
~
a(t)
t E (0,7r)
for of all
Xi(O) = Xi(7r) = TJi
~
TJi+l
(condition (3.24)),
(condition (3.25)),
(3.26) By (3.1) X~/(t) ~ supF(t,Xi(t)), that is, condition (3.23) holds. By (3.26) and (3.24) the limit x(t) = limi->oo Xi(t) ~ a(t) exists. The solution space of the system (3.5) belongs to Rc(U). The inequality x(t) ~ a(t) and (3.15) (with '"'I, pointed according to (3.21) for K = [inf a([c, d]), sup xo([c, d])]) imply that the function XI[c,d] solves (3.1). Thus the function X solves (3.1) on (0,7r). It remains to show that
limx(t) t->O
= 0,
limx(t)
t->7r
= o.
Both equalities may be proved in an analogous way. Restrict the considaration to the first one. Take an arbitrary c > O. Find an index i such that TJi < ~. Find a number 8 > 0 such that Xi(t) < TJi + ~ for t < 8. For t < 8 we have which was required. The theorem is proved.
•
CHAPTER 15
BEHAVIOR OF SOLUTIONS
In previous chapters we have developed a rather strong Cauchy problem theory. That is, we have shown how to establish the presence of the main properties of solution spaces. This allows us to give some further account of a part of the theory of ordinary differential equations at an axiomatic level. We have already made some steps in this direction. In this and in the next chapter we continue the development of the axiomatic theory of solution spaces. The classical theory of ordinary differential equations give patterns to imitate. Although we keep in the main to the general line of the account of the theory, we change essentially the account of particular questions. In the framework of the axiomatic approach many arguments of classical theory lose its validity. The work of the construction of the new theory does not remain unpaid. We have shown how we can check the fulfilment of axioms of the theory and investigate various equations. Thus all our results have a larger domain of applications than results of the classical theory. Geometric aspects of the theory of Ordinary Differential Equation become expanded, in particular, on equations with discontinuous right hand sides and on differential inclusions. For equations with continuous right hand sides we obtain also new tools of investigation. 1. Autonomous and asymptotically autonomous spaces
Let V be an open subset of the space IRn. The space Z ~ C s(IR x V) is called autonomous if it is closed with respect to translations on 1R, i.e., if Z E Z and to E IR then the function Zl defined on the set to + 7f(z) by the formula Zl(t) = z(t - to) belongs to the space Z. The set of all autonomous spaces Z E R(1R x V) is denoted by A(V). Evidently the solution space of an autonomous equation y' = f(y) (i.e., when the right hand sides does not depend on t explicitly) is autonomous. Let * denote an arbitrary set of axioms of our theory (Le., an expression as c, ce, cek and etc.). Put A.(V) = A(V) n R.(lR x V). So we define the meaning of the notation as Ace(V), Aceu(V), etc ..
442
CHAPTER 15 For Z E A(V) and M <;;; V we call diamz M = sup( {b - a: [a, b] = 7r(z), Z E Z, Im(z) <;;; M} U {O})
the diameter of the set M with respect to the space Z. For diamz M we allow the value 00 too; moreover: Lemma 1.1. Let Z E Ace(V), K be a compact subset of the set V and diamz K = 00. Then there exists a function Zo E Z+ such that Im(zo) <;;; K and 7r(zo) = [0, (0). Proof. By virtue of the equality diamz K = 00 for every i = 1,2, ... there exists a function Z; E Z such that Im(z:) <;;; K and l7l"(z*) ~ i. By virtue of the autonomy of the space Z we can assume in addition that inf 7r(z:) = O. By virtue of the compactness of the set K we can assume that the sequence {zi(O): i = 1,2, ... } converges to a point y E K. For every i = 1,2, ... fix a function Zi E Z-+ extending the function zi (such a function exists by Lemma 6.6.2). Apply Lemma 7.2.1 to the sequence of the spaces a = {Zi == Z : i = 1,2, ... }, to the sequence of the functions {Zi: i = 1,2, ... }, and to the function z*, 7r(z*) = {O}, and z*(y) = O. (Since Z E Ac(V), the sequence a converges to the space Z, see Remark 7.6.1. By virtue of the condition Z E Ae(V) the function z* belongs to Z). Let z E Z-+ denote a function existing by Lemma 7.2.1. For every segment I <;;; J = 7r( z) n [0,(0) the inclusion I <;;; [0, i] is true, beginning with some number i. Therefore I <;;; 7r(z:) <;;; 7r(Zi), Zi(I) = zi(I) <;;; K. The latter inclusion and the corresponding limit passage as i -+ 00 imply the inclusion z(I) <;;; K. In view of the arbitrariness in the choice of the segment I this implies the inclusion z(J) <;;; K. Since the set K is compact and Z E Ace(V), by Lemma 5.6.1 this is possible only when J = [0, (0). It remains to put Zo = zlJ. The lemma is proved. • Remark 1.1. Let z E Ac(V), K E expc V and diamz K ~ c > O. Then there exists a function z E Z such that Im( z) <;;; K and 7r( z) = [0, c]. This follows easily from the continuity of the mapping 7r, the compactness of the set Z[O,cjXK, the definition of diamz K, and Remark 2.3.4. Lemma 1.2. Let Z E Ac(V), {Ki: i = 1,2, ... } be a decreasing sequence of compact subsets of the set V, c > 0 and for every i = 1,2, ... , diamzKi ~ c. Then diamz(n{Ki: i = 1,2, ... }) ~ c. Proof. For every i = 1,2, ... fix a function Zi E Z such that Im(zi) <;;; K and l7l" (z;) ~ c. By virtue of the condition of the autonomy we can assume in addition that inf 7r(Zi) = O. Let zi = zil[o,cj' The graphs of the elements of the sequence {z;: i = 1,2, ... } lie in the compactum [0, c] x K. By virtue of the condition Z E Ac(V) this sequence has a subsequence converging to a function z* E Z. For it, 7r(z*) = [0, c] and Im(z*) <;;; n{Ki: i = 1,2, ... }) that gives what was required. The lemma is proved. •
443
Behavior of solutions.
Corollary. With the notation of Lemma 1.2 let diamz Ki = 00. Then diamz(n{Ki : i = 1,2, ... }) = 00. • A point y of the set V is called a stationary point of the space Z E A(V), if the constant function Z == y, 7r(z) = JR, belongs to Z-+. Evidently the diameter of a stationary point (of the space Z with respect to the space Z) is infinite. On the other hand, if the diameter of a point y is positive then the condition of the autonomy of the space Z implies that the function Z == y, 7r(z) = JR, belongs to the space Z-+, and hence the point y is stationary. If a point is not stationary then Lemma 1.2 implies that the point possesses neighborhoods of arbitrary small diameter (with respect to the space Z). A more general example of a set of the infinite diameter (with respect to the space Z) than a stationary point is the set of values of every periodic function Z E Z-+. So: Lemma 1.3. Let Z E Ace(V), a function Z E Z take in two different points of7r(z) the same value and Im(z) ~ M ~ V. Then diamz M = 00 . • Lemma 1.4. Let Z E Ace(V), Ko be a compact subset of the set V and diamz Ko = 00. Then there exists a compactum K ~ K o, diamz K = 00, which does not contain proper compact subsets of infinite diameter with respect to the space Z . Proof. Fix a countable base {3 = {Bi: i = 1,2, ... } of the compactum Ko (such a base exists by Theorems 1.6.8 and 1.6.9). Sequentially construct compacta Ko :2 Kl :2 K2 :2 K3 :2 ... of infinite diameter with respect to the space Z. Let the compacta K o, . .. ,Ki be fixed. Two cases are possible.
In this way we obtain the sequence needed. Show that the set : i = 1,2, ... } satisfies the required conditions. Its compactness follows from Theorems 1.3.4 and 1.6.1. The estimate diamz K = 00 follows from Corollary of Lemma 1.2. Now check the fulfilment of the latter condition: the diameter of every proper compact subset of the set K with respect to the space Z is finite. Assume the opposite. Let K* be a proper compact subset of the set K and diamz K* = 00. Fix an arbitrary point y E K \ K* and take an (arbitrary) element Bi of the base (3, which contains the point y and which does not meet the set K*. Then
K
= n{ K i
K i - 1 \ Bi ~ K \ Bi ~ K*, diamz(Ki_l \ B i ) ~ diamz K* =
Ki
= Ki-
1 \
B i , Y ¢ Ki(~ K),
00,
444
CHAPTER 15
which contradicts the choice of the point y. Thus our assumption is false and the lemma is proved. • Denote by B*(V) the set of all couples (Z, Zoo), where:
(1.2) Z E R(U) for some open subset U of the set JR. x V; (1.3) for every compact subset K of the set V and every number c ;:: 0 (*) if a sequence {Zi: i = 1,2, ... } ~ Z is such that Im(zi) ~ K, bi - ai ~ c, where [ai, bi ] = 71'(zd, and ai - t 00 as i - t 00, then the sequence {z;: i = 1,2, ... }, where 71'(z;) = 71'(Zi) - ai and z;(t) = zi(ai + t), has a subsequence converging to an element of Zoo' The slightly cumbersome condition (*) may be stated more simply. Let us use the notions introduced and discussed above. Denote by Z(8) the set {zo: Z E Z}, where the function ZO is defined on the set 71'(z) - 8 by the formula ZO(t) = z(8 + t). With this notation condition (1.3) is equipotent to the condition (1.4) for every sequence of numbers 8i - t 00 (as i - t 00) the sequence of the spaces {Z(8 i ): i = 1,2, ... } converges in JR. x V to the space Zoo' Thus we can use tools of our Cauchy problem theory to verify the fulfilment of condition (1.3). For a bounded set U condition (1.3) obviously holds: they have no sequence satisfying the assumptions of (1.3*). In applications, as rule, it is sufficient to consider the narrower set B(V) of all couples (Z, Zoo) satisfying conditions (1.1), (1.3) and condition (1.2) with the additional assumptions, that U = (a, 00) x V for some a E JR. U {-oo} and Z E Rce(U). The introduced notion generalizes the concept of solution space of an asymptoticaly autonomous equation and tallies its development, see [Mar], [MS], [Se1], [Se2], [LS], [Arl. Example 1.1. Let a multi-valued mapping F : V - t JR.n be upper semicontinuous and have nonempty compact convex values. Consider the mapping F(y) as a function on two arguments t and y with JR. x V as the domain of definition. The mapping does not depend explicitly on the first of these two arguments. By results and remarks of §§2.5, 6.4, 7.1, 8.2 and 8.3 the autonomous space D(F) satisfies conditions (c), (e), (k) and (n). Example 1.2. If Z E Ace(V), then (Z, Z) E B(V) (see Remark 7.6.1). Example 1.3. Let (Z, Zoo) E B*(V) and Zl ~ Z, Zl E R(JR.x V). Then (Zl' Zoo) E B*(V). If here Zl E Rce«a, 00) x V), then (Zl' Zoo) E B(V).
Behavior of solutions.
445
Example 1.4. Let a non-negative function cp be defined and be locally Lebesgue integrable on the interval (a, 00) of the real line and
f
8+C
lim
(1.5)
8-+00
cp(t)dt = 0 for every c> O.
8
Wth the notation of Example 1.1 let G(t, y) = Of(F(y), cp(t)). As in Example 1.1 the space D(G) satisfies conditions (c), (e), (k) and (n). By Theorem 8.2.2 and Remark 8.2.1 condition (1.5) imply the membership
(D(G), D(F))
E
B(V).
Notice now two conditions which are sufficient for the fulfilment of (1.5). lim cp(t) = O.
(1.6)
t-+oo
The fulfilment of (1.5) follows from (1.6) by Fatou's lemma 4.7.7.
f
00
(1.7)
cp(t)dt < 00.
a
By (1.7)
f
8+C
}~~
8
f
S+C
cp(t)dt =
}~~
a
f 8
cp(t)dt -
}~~
a
f
00
cp(t)dt =
f
00
cp(t)dt -
a
cp(t)dt = 0
a
for every c> 0, which means the fulfilment of (1.5). Example 1.5. In addition to the notation of Example 1.1 let Yo E V, Uo E IRn and for every vector u E F(yo) the scalar product (u, uo) is positive. For every t E IR and y E V put G(t, y) = F(y) + te-tIlY-Yolluo. By results and remarks of the sections mentioned in Example 1.1 we have D(G) E Rcekn(1R x V). Let U be an open subset of V. Let the closure of U be compact, lie in V and not contain the point Yo. The fulfilment of the condition (D(G, IR xU), D(F, IR xU)) E B(U) follows easily from Theorem 8.2.2 and from Remark 8.2.1. Nw let c = ~ inf{(u, uo) : u E F(yo)} ( > 0 in view of our assumptions), L = {u: u E IRn , (u, uo) > c}. By virtue of the upper semicontinuity of the mapping F the set Vo = {y: y E V, F(y) ~ L} is open. For t E IR and u E Vo put h(t, u) = (u, uo). Theorem 4.10.1 implies that the spaces D(G,1R x Vo) and D(F,1R x Va) lie in M1(h,c) (the notation from §7.4). Theorem 7.6.2 with W = Ml (h, c) implies the fulfilment of the condition
(D(G, IR x Va), D(F, IR x Va)} E B(Va).
CHAPTER 15
446
By Theorem 7.6.4 (D(G), D(F)) E B(V).
2. Limit sets Let a continuous function z be defined on an interval (a, b) or on a half interval [a, b), a < b, of the real line and take values in the space ~n. Put
O+(z) = n{[z([t, b))]: t E 7r(z)}. Respectively, if a continuous function z is defined on a interval (a, b) or on a half interval (a, b], a < b, put
O-(z)
= n{[z((a, t])]: t
E 7r(z)}.
We mean these sets in the title of the section. We can define them in another way too. The reader may prove easily that:
O+(z) = limtopsupz(t), t->b
O-(z)
= lim topsupz(t). t->a
Evidently properties of these sets are similar. Assertions of this section will be stated for one of them only. Lemma 2.1. Let a continuous function z be defined on an interval (a, b) or on a half interval [a, b), a < b, of the real line and take values in the space ~n. Let the set 0+ (z) be nonempty and compact. Then for every point t E 7r(z) the set z([t, b)) U O+(z) is compact. Proof. Assume the opposite. By virtue of the compactness of the set O+(z) this means the existence of points Sk E [t, b) (for k = 1,2, ... ) such that the sequence {Z(Sk): k = 1,2, ... } has no limit points in the space ~n. In view of the continuity of the function z this is possible only if Sk - t b. By the definition of the set O+(z) we have IIZ(Sk)lI-t 00 and we can assume in addition that all points z(sd, k = 1,2, ... , lie outside the set O(O+(z),2) (in the opposite case we pass to a corresponding subsequence). Since b is greater than every element of the sequence {s k: k = 1, 2, ... }, we can assume in addition that Sl < S2 < S3 < . .. (here too, in the opposite case we pass to a corresponding subsequence). By virtue of the connectedness of the set [Z([Sk' b))] (see Example 2.6.4, Corollary 1 of Theorem 2.6.4 and Theorem 2.6.2) and by the definition of O+(z) Corollary 1 of Lemma 2.6.1 implies easily the nonemptiness of the set
Mk = [Z([Sk' b))] n (O,(O+(z), 2) \ O(O+(z), 1)) = Z([Sk' b)) n (O,(O+(z), 2) \ O(O+(z), 1)). So we obtain a non-increasing sequence {Mk : k = 1,2, ... } of nonempty bounded (see Theorem 2.4.1 and Remark 2.3.3) closed (hence, compact)
Behavior of solutions.
447
sets. By the Corollary of Theorem 1.6.3 the set M = n{Mk: k = 1,2, ... } is nonempty. By the definition of O+(z) the set M lies in O+(z). By its construction the set M lies outside the neighborhood O(O+(z), 1) of the set O+(z). The obtained contradiction shows that our assumption is false. The lemma is proved. • Theorem 1.7.2 allows us to obtain from the proved lemma Corollary. Under the hypotheses of Lemma 2.1 values of the function z(s) converge to the set O+(z) in the sense of condition (1.7.1) as s ~ b.• In the proof of Lemma 2.1 we have already noticed the connectedness of sets [z([t, b))l (with the notation of the proof they were [Z([Sk, b))]). Therefore Lemma 2.1 and the Corollary of Theorem 3.3.3 imply: Lemma 2.2. Under hypotheses of Lemma 2.1 the set O+(z) is con-
.
~~
We will use the proved assertions later. Let us return to the topic of the previous section. As there let V denote an open subset of the space Rn. Lemma 2.3. Let (Z, Zoo) E B*(V), {Zi: i = 1,2, ... } ~ Z, lim topsUPi ..... oo Im(zi) ~ K E expc V, 7r(Zi) = [ai, bil (i = 1,2, ... ), ai --? 00 and y E lim top supzi(ai) (respectively, y E limtopsupzi(bi)). Then: i--+oo
i-+oo
if numbers bi - ai, i = 1,2, ... , are bounded in totality then the sequence of the functions {z;: i = 1,2, ... }, where 7r(z;) = 7r(Zi) - ai and z;(t) = zi(ai + t) for i = 1,2, ... , contains a subsequence converging to a function z* E Zoo such that z*(inf 7r(z*)) = y (respectively, such that z*(SUP7r(z*)) = y), if bi - ai ~ 00 (as i ~ 00) then there exists a function Z E Z! (respectively, a function Z E Z~) such that 7r(z) = [0,00) (respectively, 7r(z) = (-00,0]), z(O) = y and Im(z) ~ K. Proof. An obvious passage to a subsequence reduces the assertion to the case, when y = limi ..... oo zi(ai) (respectively, y = limi ..... oo zi(b i )). If numbers bi - ai, i = 1,2, ... , are bounded in totality, we then obtain what was required from the comparison of the hypotheses of the lemma with condition (1.3*). Now consider the case where bi - ai --? 00 as i --? 00. Take an arbitrary c> 0 and denote A = {i: i = 1,2, ... , bi - ai ~ c}. Condition (1.3*) in relation with the sequence {zil[ai,ai+ c ]: i E A} (respectively, {zil[bi-C,bi] : i E A}) gives the existence of a function z; E Zoo with lll"(zJ = c, Im(z;) ~ K and z;(inf 7r(z;)) = y (respectively, z;(sup 7r(z;)) = y). The condition of the autonomy of the space Zoo gives the possibility assuming in addition that inf7r(z;) = 0 (respectively, SUp7r(z;) = 0). For c = 1,2, ... fix an arbitrary extention z~* E Z~+ of the function z~ (see Lemma 6.6.2). Apply Theorem 7.2.2 to the sequence of the spaces a = {Zc == Zoo: C = 1,2, ... }, the sequence of the functions {z~*: c = 1,2, ... } and the sequence ofthe points {teO == 0: c = 1,2, ... } (by Remark 7.6.1 the sequence a
448
CHAPTER 15
converges to the space Zoo). Let z** be a function existing by Theorem 7.2.2. If a segment I lies in the half interval J = [0,00) n 7r(z**) (respectively, in the half interval J = (-00, O]n7r(z**)), then I ~ 7r(z;), z;*(I) = z;(I) ~ K, beginning with some c = 1,2, .... Therefore z**(I) ~ K. In view of the arbitrariness in the choice of the segment I this means that z**(J) ~ K. The compactness of the set K and Lemma 6.6.1 imply that J = [0,00) (respectively, J = (-00,0]). It remains to put z = z**IJ. The lemma is proved. • Theorem 2.1. Let (Z,Zoo) E B*(V), z E Z+, 7r(z) = [a,oo) and y E V n n+(z). Then there exists a function z* E Z~+ such that y E Im(z*) ~ n+(z). If here the set n+(z) is compact and lies in V then
we can require, in addition, that 7r(z*) = (-00,00). Proof. Fix a sequence of points {ai: i = 1,2, ... } ~ 7r(z), a < al < a2 < a3 < ... , ai
-t
00, z(ai)
-t
y.
By condition (1.4) the sequence of the spaces a = {Zi = Z(ai) : i = 1,2, ... } converges in every region of the form (to, 00) x V, to E JR, to the space Zoo. Let Zo E Z-+ be an extention of the function z in the left. For i = 1,2, ... and t E 7r(zo) - ai put Zi(t) = zo(t + ai). Evidently Zi E Zi-+. When we apply Theorem 7.2.2 to the sequence of the spaces a, to the sequence of the functions {Zi: i = 1,2, ... } and to the sequence of the points {t i == 0: i = 1,2, ... } we obtain the existence of a function z* E Z~+ with z*(O) = limi ..... oo Zi(O) = y. Let [a', b/] ~ 7r(z*). Since zi([a' , b']) = zO([ai + a', ai + b']), we have z*([a' , b']) ~ n+(z). In view of the arbitrariness in the choice of the segment [a', b'] this means, that Im( z*) ~ n+ (z). The first part of the theorem is proved. To prove the second part of the theorem choose the sequence {ai : i = 1,2, ... } in order to fulfill the additional condition
ai+l - ai
~
i for every i = 1,2, ....
When we apply Lemma 2.3 to the sequence of the functions {zl[ai,ai+l] : i = 1,2, ... } we obtain (in view of Lemma 2.1) the existence of functions z; E Z~ and z; E Z! such that 7r(z;) = (-00,0], 7r(z;) = [0,00), z;(O) = z;(O) = y and Im(zd U Im(z2) ~ n+(z). It remains to put
z(t) = {z;(t) z;(t)
for t E (-00,0], for t E [0,00).
The theorem is proved. • Corollary. Let Z E Ace(V), z E Z+, 7r(z) = [a,oo) and y E V n n+(z). Then there exists a function z* E Z-+ such that y E Im(z*) ~ n+(z). If
Behavior of solutions.
449
the set n+{z) is compact and lies in V then we can require in addition that
7r{z*) = (-oo, 00) (in this case diamz n+{z) = 00). To obtain the corollary it is sufficient to refer to remarks of Example 1.2. ..
3. Some geometric properties of solution spaces In this section we generalize Theorems 1.1 and 1.2 of Chapter VII of [Ha]. Let V be an open subset of the space ~n. Theorem 3.1. Let Z E Ace{V), Vo be an open subset of the set V, H be a closed (in Vo) subset of the set Vo. Assume that the set Ov H is not compact and may be represented as the union of two disjoint sets HI and H 2 ; moreover, the set HI is compact, and for every point y of the set H 2 :
(3.1) there exists a function z E Z such that 7r(z) = [c,O] for some c z(O) = y, z(c) E Hand Im(z) ~ [Vo].
< 0,
Then there exists a function Zo E Z-UZ+ such that Im(zo)n(HI UH2 )"I 0 and Im(zo) ~ [Vo]. (Here [M] = [M] v denotes the closure of the set M in V.) Proof. I. Let y be an arbitrary point of the set H 2 • Compare two
conditions: (3.2) there exists a function z E Z such that z{inf7r{z)) E HI, Im(z) and z{sup 7r{z» = y;
~
[Vo]
and (3.3) there exists a function z E Z- such that z{sup 7r{z» Im{z) ~ [Vo].
= y and
II. Assume that for a point y of the set H2 condition (3.2) is not fulfilled. Denote by M the set of all functions z E Z satisfying the conditions SUP7r(z) = 0, z(O) = y and Im(z) ~ [Vo]. For z E M put a(z) = inf{t: t E 7r(z), z(t) E H}. Let a E a(M). Then a = inf7r(z) for a function z EM. If z{a) E (H), then a > b, where b = inf a{M) by virtue of the condition Z E Ae{V). If z{a) E ovH, then by our assumption z{a) rt. HI' Therefore z{a) E H 2 • By (3.1) a > b. Thus a{M) ~ (b, OJ, where b = inf a{M) < O. Fix a sequence of points {b k : k = 1,2, ... } ~ a{M), bk --+ b, and
a sequence of functions {Xk: k = 1,2, ... } function Zk E Z-+ extend the function Xk'
~
M, [b k , OJ
=
7l"{Xk)'
Let a
450
CHAPTER 15
Apply Theorem 7.2.2 to the sequence of the spaces {Zk == Z : k = 1,2, ... }, to the sequence of the functions {Zk: k = 1,2, ... } and to the sequence of the points {tk == 0: k = 1,2, ... }. Let z* denote a function existing by Theorem 7.2.2 (note: z*(O) = y). Let the segment [c,O], c < 0, lie in its domain of definition. If c > b, then the convergence in Theorem 7.2.2 implies the inclusion z*([c, 0]) ~ [Vo]. If b E 7r(z*) here then the limit passage as c --t b proves the inclusion z*([b, 0]) ~ [Vo]. Here z(bd E [H] and bk --t b. By Theorem 3.5.4 this implies that z*(b) E [H] and bE a(M). This contradicts the established inclusion a(M) ~ (b,O]. Let 7r(z*) ~ (b, 00). Here z*([c, 0]) ~ [Vo] for every number c from the set B = 7r(z*) n (-00,0]. This implies that Zo E Z-, SUP7r(zo) = 0, zo(O) = Y, Im(zo) ~ [Vo] for Zo = Z*IB' This means the fulfilment of (3.3). II. Assume now that condition (3.2) holds for every point Y E H 2 • Denote by M the set of all functions Z E Z satisfying the conditions: Im(z) ~ [Vo], inf7r(z) = 0, z(O) E HI and z(SUp7r(z)) E H 2 • Let the mapping ~ : M --t H2 be defined by the formula ~(z) = z(SUp7r(z)). By Theorems 3.5.1 and 3.5.4 the mapping ~ is continuous. By our assumptions ~(M) = H 2 • Theorem 1.7.8 implies the noncompactness of the set M. Let a = {Zk: k = 1,2, ... } be a sequence of elements of the set M without limit points in M. By virtue of the compactness of the set HI we can assume in addition that the sequence {Zk(O): k = 1,2, ... } converges to a point Yo E HI' For every k = 1,2, ... fix an extension z;; E Z-+ of the function Zk and apply Theorem 7.2.2 as in 1. Let z* denote a function existing by Theorem 7.2.2 (note: E 7r(z*) and z*(O) = Yo). Let {z;;: k E A} be a corresponding subsequence. For every c from the set J = 7r(z*) n [0,00) we have [0, c] ~ 7r(z;;), beginning with some k = ko E A. Show that the set Al = {k : k E A, k ~ ko, sup7r(zd ~ c} is finite. Assume the opposite. By virtue of the compactness of the segment [0, c] the sequence {sup 7r(zd: k E Ad has a subsequence {SUP7r(Zk): k E A 2 } converging to a point d E [O,c]. By Theorem 3.5.4 the sequence {Zk: k E A 2} converges to the function Z*I[O,d] E M, that contradicts the choice of the sequence a. Thus z*([O, c]) ~ [Vo]. In view of the arbitrariness of c E J this means that the function Zo = z*IJ E Z+ satisfies the imposed condition. The theorem is proved. • Theorem 3.2. Let all hypotheses of Theorem 3.1 hold, Z E Ak(V), the set H be connected and the set HI be nonempty. Assume that:
°
°
(3.4) for every function Z E Z- satisfying the conditions E 7r(z) and z(O) E H2 there exists a number c E (-00,0] n7r(z) such that z([c, 0]) ~ [Vo] and z(c) (i [H].
Behavior of solutions.
451
Then there exists a function Zo E Z- U Z+ such that Im(zo) n HI 1= 0 and Im(zo) ~ [Vo]. Proof. Put MI = U{Im(z): z E Z-, Im(z) ~ [Vo]} and M2 = U{Im(z) : z E Z, z(inf 7r(z)) E HI, Im(z) ~ [Vo]}. I. Let the set M2 be noncompact. There exists a sequence {Yk : k = 1,2, ... } ~ M2 without limit points in M 2. For every k = 1,2, ... fix a function zZ E Z-+ such that 0 E 7r(zZ),zZ(O) E HI and zZ(td = Yk for some tk E 7r(zZ) n [0, (0). Now repeat the reasoning of step II of the proof of the previous theorem (with Zk = zZI[o,tkj). We obtain the existence of the needed function. II. If the set M2 is compact and if diamz M2 = 00 then our assertion follows easily from Lemma 1.1. III. It remains to consider the last possibility: the set M2 is compact and diamz M2 < 00. Theorem 7.2.2 implies easily the closed ness of the set MI. Take an arbitrary point Y E [H] and consider an arbitrary function z E Z-, z(SUp7r(z)) = y. If the set Im(z) n HI is nonempty then Y E M 2 • If the set Im(z) n HI is empty then (3.4) implies the inclusion Im(z) ~ [Vo]. So Y E MI. Thus [H] ~ MI U M2 • The set M2 n [H] is nonempty because it contains the nonempty set HI. Since it is compact it cannot contain the noncompact closed set H 2 • The set H is connected. Therefore its closure [H] is connected too. Remarks of §2.6 and mentioned facts imply the nonemptiness of the set M = MI n M 2 • Being the intersection of a compactum and of a closed set the set M is compact. We have: diamz M ~ diamz M2 < 00. Let a = diamz M (so a < (0). The compactness of the set ZI[O,ajxM implies the existence of a function z E Z such that 7r(z) = [0, a] and Im(z) ~ M. The membership z(O) E MI implies the existence of a function ZI E Zsatisfying the conditions sup 7r(zd = 0, ZI (0) = z(O) and Im(zd ~ [Vo]. Let b = inf 7r(z). If z(O) E HI then the function Zo = ZI satisfies the imposed conditions. In the opposite case fix a function Z2 E Z such that 7r(Z2) = [e,O] for some c < 0, Z2(0) = z(O), Im(z) ~ [Vo] and z2(e) E HI. The membership Z E Acek (V) implies the existence of a number c; > 0 such that for It I < c; the set St = {z*(t): z* E Z, 7r(z*) :3 0, t, z*(O) = z(O)} is connected. Let 'Y E (max{b, e, -c;}, 0). The set S,,! contains the points Zl (r) E Ml and Z2 (r) E M 2. By virtue of the connectedness of the set S'"( this implies the existence of a point Y E S'"( n Ml n M 2 • Hence for a function Z3 E Z we have 7r(Z3) = ['Y, 0], Z3(r) = Y and Z3(0) = z(O). Now define the function for t E [0, aj, for t E ['Y, OJ,
452
CHAPTER 15
7f(Z4) = [" a]. Since 1... (z4) > a and Im(z4) ~ M, the existence of Z4 contradicts the definition of the number a. The theorem is proved. •
4. Periodic spaces Let U be an open subset of the product IR x IRn. Let a be a positive number. A space Z ~ Cs(U) is called periodic with period a, if for every Z E Z and k = ±1, ±2, ... the function Zl (t) 7f(zd = 7f(z) + ka, belongs to the space Z.
=
z(t - ka),
The set of all periodic spaces Z E R(U) with period a is denoted by pa(u). As in the case of autonomous spaces, for every set * of axioms of our theory p.a(u) = pa(u) n R.(U). Notice that here we speak about the periodicity of a space on the whole and we do not mean the periodicity of its elements. If a space Z E pa(u) contains all functions with one point domain and with graphs lying in U (for instance, if Z E P;(U) or Z E Pea(U)) then the set U possesses the corresponding periodicity too. As for autonomous spaces, here we can introduce an object corresponding to a perturbed case. But such considerations remain outside the line of our account. If the right hand side of an inclusion y' E F( t, y) is periodic with a period a in t (i.e., for every (t,y) E U and k = ±1,±2, ... we have (t+ka,y) E U and F(t + ka, y) = F(t, y», then D(F) E pa(u). This remark gives main patterns of spaces from pa(u). In this and in the next sections we restrict the discussion to one question only, although it is rather interesting. Notice first: Lemma 4.1. Let a bounded closed subset H of the space IRn be homeomorphic to the closed ball, -00 < tl < t2 < 00, ttl, t 2] X H ~ U, Z E Rceu(U). Assume that:
(4.1) for every t E ttl, t 2) and y E H there exists a number CI > 0 such that if C E (0, cd, Z E Z, 7f(z) = [t, t + c] and z(t) = y, then Im(z) ~ H. Then there exists a function Zl E Z such that 7f(zd = [tl' t2], Im(zd ~ Hand Zl(t 1 ) = ZI(t 2). Notice that under the additional assumption y E (H) condition (4.1) holds obviously. So it remains valued only for points y E 8H. Proof. I. Associate to every point y E H a function Zy E Z+ such that tl = inf 7f( Zy) and Zy (td = y. By virtue of condition (u) such a function is unique.
Behavior of solutions.
453
II. With the notation of! let a(y) = sup{t: t E 7r(z), Z([tl' t 2]) ~ H}. By (4.1) we have a(y) > t l . Show that a(y) ~ t 2. Assume the opposite, i.e., that a(y) < t 2 • The graph of the function Zy lies in the compactum [tl' t2J x H. By condition (c) the sequence
has a subsequence converging to a function z* E Z. Evidently the function z* extends the function Zy and SUP7r(z*) = a(y). Now (4.1) and the membership Z E Rce(U) imply the possibility of extending the function z* on the right over a(y) to a function z** E Z with the fulfilment of the condition Im(z**) ~ H. This contradicts the definition of a(y). III. By Lemma 8.1.1 II imply the compactness of the set q, = {z : z E Z, 7r(z) = [tl, t 2], z(t l ) E H}. By Theorems 2.2.1 and 3.5.4 and by the condition Z E Ru(U) this implies the continuity of the mapping 6: H ~ H, 6(y) = Zy(t2). Theorem 2.8.4 now implies, that Zy(t2) = Y = Zy(t l ) for some y E H. Put Zl = Zy. The lemma is proved. • By virtue of thre possibility of extending by periodicity the functions Lemma 4.1 implies: Theorem 4.1. Let a subset H of the space ~n be homeomorphic to the closed ball, ~ x H ~ U, Z E Pco."u(U). Assume that
(4.2) for every t E ~ and y E H there exists a number CI > 0 such that if Z E Z, 7r(z) = [t, t + cJ and z(t) = y, then Im(z) ~ H.
c E (0, cd,
Then there exists a periodic function Im(zd ~ H.
Zl
E Z-+ of the period a such that •
5. Limit passages in the space Rc(U) Quite often we have discussed the following construction. In order to prove the fulfilment of a condition for solutions of an equation y' = f(t, y) we approximate the right hand side f by functions fk' k = 1,2, ... such that for the equations y' = !k (t, y) the corresponding condition holds and then we check its keeping under the limit passage fk ~ f. Let us discuss Theorem 4.1 from this point of view. It is simpler not to watch for the periodicity of spaces under consideration and to carry out the argument on the level of Lemma 4.1. Keep the notation of the previous section. Lemma 5.1. Let H be a compact subset of the space ~n. Let -00 < tl < t2 < 00 and [tl' t2J x H ~ U. Then the set of spaces Z E Rc(U) satisfying the condition
454
CHAPTER 15
(5.1) there exists a function z E Z such that 7r(z) z(t l ) = z(t 2),
= [tl' t 2],
Im(z) ~ Hand
is closed in the space Rc(U). Proof. The proof is left to the reader as an easy exercise. Consider the following situation.
•
A. A bounded closed subset H of the space ~n is homeomorphic to the closed ball, R x H ~ U and -00 < tl < t2 < 00. Lemma 5.2. Let condition A hold. Let a space sequence {Zk : k = 1,2, ... } ~ Rceu(U) converge in U to a space Z E Rc(U). Assume that
(5.2) if zZ E Z;+, tl E 7r(zZ), zZ(td E Hand OH is a neighborhood of the set H, then [tl, t2J ~ 7r(zZ) and zZ([t l , t 2]) ~ OH, beginning with some k = ko .
Then condition (5.1) holds. Proof. 1. The set H is an absolute retract. Let r : ~n - t H denote a retraction, G6 = {x: x E ~n, IIx - r(x)1I < o}. II. Assume that our assertion is false. Then for some c > 0
III. For k = 1,2, ... and h E H denote by Zkh an unique function z E Zi:+, satisfying the conditions tl E 7r(z) and z(t l ) = h. Let ko be pointed for OH = Go according to (5.2). From II if hE H, then IIZkh(td - Zkh(t 2) II > 2c, beginning with some k = kl ~ k o. . For k = kl consider the mapping f : H continuous and
-t
H, f(h) = r(zkh(t 2)). It is
for h E H. The last estimate contradicts Theorem 2.8.4. The lemma is proved. • Lemma 5.3. Let a sequence a = {Zk: k = 1,2, ... } ~ Rceu(U) converge in U to the space Z E Rc(U). Let conditions A and (4.2) hold. Then the sequence a satisfies (5.2), beginning with some k = ko. Proof. The proof reduces to referring to Theorem 7.2.2 and to Lemma 5.1.
•
Behavior of solutions.
455
The last lemma immediately implies Theorem 5.1. Let Z E [Rceu(U)]Rc(U)' Let conditions A and (4.2) hold. • Then condition (5.1) holds too. Consider the following situation. B. Condition A holds, F : U - IRn is a multi-valued mapping, a (single valued) function'ljJ : IRn - IR is continuous and has continuous derivatives, (H) = {x: x E IRn , 'ljJ(x) < OJ, grad'ljJ(x) 0 for x E 8H, for every e > 0 there exists a neighborhood V of the set 8H such that ~v ~ (-00, e], where ~M = {(u, grad 'ljJ(y)): u E F{t, y), tl ~ t ~ t 2 , Y E M, (t, y) E U}.
t=
Start with the following auxiliary consideration. C. Let in addition to B p < 0, IR x W neighborhood W of the set 8H.
~
U and ~w ~ (-oo,p) for a
By Lemma 10.2.1 for every function Z E D{F, IR x W) the function 'ljJ{z{t)) is generalized absolutely continuous and ('ljJ{z{t)))~ < p for almost all t E 7r{z). This implies easily that the space Z = D{F) satisfies condition (4.2).
Theorem 5.2. Let under the assumptions of B the mapping F satisfy locally the Davy conditions. Then the space Z = D{F) satisfies condition (5.2). Proof. I. By Lemma 2.4.2 for some e > 0 the set 0f([t 1 ,t2 ] x H,e) is compact and lies in U. Therefore we can assume in addition that F E Qd(cp){U) for a locally Lebesgue integrable function cp (if it is necessary we pass to the set O([t 1 , t 2 ] x H, e) instead of U) Likewise we can assume that sup II grad 'ljJ(U) II ~ m for some m > O. II. Assume in addition the fulfilment of condition C and notice that the fulfilment of the condition D{F) E Rc{U) already follows from other our assumptions. In §8.2 we have constructed a sequence {Ii: i = 1,2, ... } of functions approximating F. Recall its properies. They are defined on the product IR x IRn , satisfy the Caratheodory conditions, and
(5.3) D(fi)
E
Rceu(1R x IRn) for every i = 1,2, ... ,
(5.4) the sequence {D(fi' U): i = 1,2, ... } converges in U to the space D{F).
By C
CHAPTER 15
456
beginning with some i. This implies the fulfilment of condition (4.1). By I and Lemma 4.1 the space (Z =) D(h U) satisfies condition (5.1). By Theorem 5.1 the space (Z =) D(F) satisfies condition (5.1) too. III. Pass to the general case. For every c > 0 the mapping
Ff:(t, y)
= F(t, y) + c grad 'IjJ(y)
satisfies the assumptions of II (by the last condition in C and by Lemma 2.4.2 the pointing of a suitable W is not difficult). The mapping
Ff:*(t, y)
= F(t, y) + [0, c] grad 'IjJ(y)
satisfies the Davy conditions. As for C1 < C2, (t, y) E U we have the inclusion F(t,y) ~ Ff:*l(t,y) ~ Ff:*2(t,y) and F(t,y) = n{Ff:*(t,y): c > O}, then Lemmas 8.2.2 and 8.2.3 imply the convergence of the generalized sequence {D(Ff:(t,y): c > O} as c --+ 0 to D(F) in the space Rc(U). By II and Theorem 5.1 we obtain what was required. The theorem is proved. • Corollary. Let condition A hold with t1 = 0 and t2 = a. Let the mapping F be periodic in t with the period a and satisfy locally the Davy conditions. Then the space Z = D(F) satisfies condition:
(5.3) there exists a periodic function Im(zd ~ H.
Zl E
Z-+ of the period a such that
•
In the Corollary of Theorem 5.2 we repeat with small modifications and expansions Theorem 3 of §14 of [Faf]. Although we follow in the main the plan of the proof of [Faf], we use some general arguments of our account too. These arguments may act in other situations. In previous chapters we have pointed to suitable limit passages in the space Rc(U). So Theorem 5.3. Let condition A hold, F E Q**(U), D(F) E Rcn(U), Fo E Q**P (U). Let a Lebesgue integrable
1) F(t, y)
n 0 / (0,
2) either Fo(t, y)
= F(t, y) or the set Fo(t, y) be nonempty and compact.
Then the space (Z =)D(Fo) satisfies condition (5.1). Proof. The proof repeats the proof of Lemma 8.7.6 with the addition of referring to Theorems 5.2 and 5.1. (If with the notation of Theorem 5.3 F(t, y) = 0 / (0,
Behavior of solutions.
457
Theorem 5.4. Let condition A hold. Let for every element V of a family 'Y of open subsets of the set U the mapping Flv satisfy Davy conditions. Let the set U \ (U'Y) be at most countable. Then the space D(F) (= Z) satisfies (5.1). Proof. The proof repeats the proof of Theorem 5.2 with the change of referring to §8.7 by referring to Theorem 10.6.1 (with i = 1 and hI (t, y) == t) or to Example 10.6.1. • Each of the last two theorems can added by a corollary similar to Corollary of the Theorem 5.2. 6. First approximation. Asymptotic stability of a stationary point In this section we show what interpretation the idea of the 'first approximation' obtains in the framework of our theory. To make notation simpler assume that the system of coordinates is such that the stationary point in question is the zero point 0 = (0, ... ,0) of the space JRn, V = 0 0 0 is a O-neighborhood of the point 0 in the space JRn (0) 0), -00 ~ a < b ~ 00, U = (a, b) x V and UI = (a, b) x JRn. A space Zo E Ri(Ud is called conic (at the point 0, which will not be mentioned in what follows) if (6.1) the function
Zo
== 0 belongs to Zo,
(6.2) for every A > 0 the change of variables cp>.{t,y) = (t, AY) transforms the space Zo onto itself. The solution space of a linear homogeneous equation is an example of a conic space. For M ~ U1 and A > 0 denote M(A) = {(t,AY): (t,y) EM}. For Z ~ Cs(U) and A > 0 denote Z(A) = (h(Z) (~ Cs(U(A»). We say that a conic space Zo E Ri(Ud is the 'first approximation' of a space Z E Ri(U), if (Z(A))U ~ (Zo)u as A ~ 00. See also Theorem 10.6.2. Example 6.1. Let Zo be the solution space of a linear homogeneous equation y' = Ay with a constant matrix A. Let Z be the solution space of an equation y' = Ay + f(t, y), where the function f is continuous and sup {
IIf(t, y)1I
lIyll
:
t E
(a, b), 0 <
lIyll
~ r
}
~ 0
as r ~ 0 (the usual condition of the 'first approximation'). Then the space Zo is the 'first approximation' of the space Z in our sense. In order to prove it, notice that for IIYII ~ ro we have the following situation. Denote Jl.(r)
= sup { III(t, lIylly)1I :
t E (a, b),
lIyll
~ r
}
.
458
CHAPTER 15
We have f.L{r)
---t
0 as r
---t
O. By putting r = r;e-, we obtain:
as A ---t 00. The last passage to the limit implies the convergence needed by theorems of the classical theory (see §7.1). Now assume that:
(6.3)
b=
(6.4)
Zo E A~(I~ x ~);
00;
and that the point {} is asymptoticaly stable with respect to the space Zo, namely (as the space Zo is autonomous) (6.5) for every number c > 0 there exists a number 0 > 0 such that if E Zo, to = inf7r(z) and IIz{to)11 ~ 0, then Ilz{t)II ~ c for every t E 7r{z);
Z
(6.6) there exists a number 00 > 0 such that for every number c > 0 there exists a number", > 0 such that if Z E Zo, to = inf7r{z), Ilz(to)11 ~ 0o, t E 7r(z) and t ~ to then Ilzo(t)11 ~ c.
+""
zt
Assertion 6.1. Under the assumptions of (6.4 - 5), if z E then SUP7r{z) = 00. Proof. We obtain what is required from Lemma 6.6.1. • Assertion 6.2. Under the assumptions of (6.4), (6.1 - 2), (6.5 - 6) there exists a number",o > 0 such that if Z E Zo and s, t E 7r(z), t ~ s + "'0, then Ilz{t)II ~ IIz~)II. Proof. Fix 00 according to (6.6). Find", according to (6.6) for c = ~. When we put "'0 = ", we obtain what was required from the conicity of the • space Zo. The assertion is proved. For T ~ 0 consider the homeomorphism 'I/J : IR x IRn ---t IR x IRn , 'l/Jr (t, y) = (t - T, y) (translation along the time axis IR). In order to include in the consideration the non-autonomous case enforce the above condition of the 'first approximation' upon the condition (6.7) (-¢r{Z{A))u) ---t {Zo)u as A ---t 00 uniformly in T {i.e., we have the convergence (-¢r; (Z (Ai)) u) ---t (Zo) u for every sequences Ai ---t 00 and Ti, i = 1,2, ... ).
Theorem 6.1. Let conditions (6.1 - 7) hold, Z E Rcr(U). Then:
459
Behavior of solutions.
(6.8) for every number E > 0 there exists a number 0 > 0 such that if z E Z, = inf7r(z) and Ilz(to)11 :s; 0 then Ilz(t)11 :s; E for every t E 7r(z);
to
(6.9) there exists a number 01 > 0 such that for every number E > 0 there exists a number 'TJ > 0 such that if z E Z, to = inf7r(z), Ilz(to)11 :s; 01, t E 7r(z) and t ~ to + 7], then Ilzo(t)11 :s; E.
Proof. 1. Prove the fulfilment of (6.8). Assume the opposite. Then for some El > 0 for every k = 1,2, ... there exists a function Zk E Z, 7r(zd = [ak,b k ]' such that Ilzdak)11 :s; ~, Ih(bdll ~ E1' When we pass (if it is necessary) to a smaller domain we achieve the fulfilment of the additional condition ~ < Ilzdt)11 < El for t E (ak, bk ). Let E = 0o, where 0o is chosen according to (6.6). Fix a number 0 < E according to (6.5). For () from the definition of V and for the 0 mentionned fix an arbitrary number 1/ E (0, min{ (), o}). Next, for k = 1,2, ... put Ak = 2kl/ and Tk = ak' For k = 1,2, ... , k > E.., select points Ck E [ak' bk ] EJ 1/ such that Ilzdck)11 = T! and ~ < Ilzk(t)11 < T! for t E (ak,d. Two (or both at once) situations are possible.
t
A. For some C> 0 the set {k: k = 1,2, ... ,
Ck -
B. For an infinite subset A of the set N
(Ck -
lim
kEA k--oo
ak
:s; C} is infinite.
ak)
=
00.
By (6.7) in the case A Assertion 10.4.1r implies the existence of a function z E Zo, such that Ilz(inf7r(z»11 = 1/, Ilz(sUp7r(z»11 = 2E and 1/ :s; II z( t) II :s; 21/ for t E 7r( z). By our choice of 0 and 1/ this contradicts (6.5). In the case B by (6.7) the same assertion 10.4.1r implies the existence of a function z E Zd such that Ilz(inf7r(z»11 = 1/ and 1/ :s; Ilz(t)11 :s; 2E for t E 7r(z). By our choice of 0 and 1/ this contradicts (6.6). Thus our assumption is false and we have what was required. II. For E = ~ select a number 01 = 0 < E according to (6.8). By virtue of the condition Z E Rcr(U) the choice of the number 01 implies that if z E Z+ and Ilz(inf7r(z»11 :s; 01 , then SUP7r(z) = 00 and Ilz(t)11 :s; ~ for every t E 7r(z).
Choose the number 'TJo according to Assertion 6.2. Prove the fulfilment of the condition (6.10) there exists O2 E (O,od such that if z E Z, z(inf 7r(z» and inf7r(z) + 'TJo :s; t:s; inf7r(z) + 27]0, then IIz(t)11
<
Ilz(inf7r(z»II. 2
:s; 02, t
E 7r(z)
460
CHAPTER 15
Assume the opposite. We obtain the existence of a sequence of functions {Zk: k = 1,2, ... }, 1I"(zd = [ak, bk], satisfying the condition
When we pass to the limit according to (6.7) and Assertion 10.4.1r with Ak = IIZ(?klll and 7k = ak, we obtain either the existence of a function Z E zit with sup 11"( z) ~ inf 11"( z) + 21/0, or the existence of a function Z E Zo with SUP1l"(z) ~ inf1l"(z) + 21/0 and II Z (sup 11" ())II Z
~
Ilz(inf21I"(Z)) II .
The first case contradicts Assertion 6.1. The second case contradicts the choice of the number 1/0 in Assertion 6.2. The contradiction obtained proves the validity of (6.10). Now (6.10) implies the fulfilment of (6.9) in an obvious way. The theorem is proved. • Example 6.2. By remarks of Example 7.5.4 the solution space of the system (6.11)
{
X'
= -x+y,
y' = -x - y.
is the 'first approximation' of the solution space of the system (6.12) The origin of coordinates is asymptoticaly stable with respect to the solution space of the system (6.11). This is easy to check by well known classical methods. By Theorem 6.1 the origin of coordinates is asymptoticaly stable with respect to the solution space of the system (6.12) too. Notice the singularity in the last terms on the line y = 0: ~ -+ ±oo as x -+ O. 7. Asymptotic estimates In this section we continue the investigation of the previous section and we obtain corresponding assertions on asymptotic behavior of solutions of equations similar to Theorem X.11.2 of [Hal. For >',7 E IR and (t,y) E IR x IRn put CPT>.(t,y) = (t - 7,>.-ly). The mapping CPT>' is obviously a change of variables in every space Y ~ C.(U). Assume that:
461
Behavior of solutions.
(7.1) U is an open subset of the product ~ ~ X ~n, V = U \ (~ x 0); (7.2) Z E R~{U), Z E Z+, 7r{z) t E 7r{z), Ilz{t)11 =F OJ;
=
X
~n, a E ~, [a,oo) x
{OJ
=
T
[to,oo) and SUpT
00,
where
~ U,
= {t:
(7.3) an open subset G of the extended line ~* = ~ U {-oo, oo} contains the set E of all limit (in ~*) values of the expression t-Iln Ilz{t)11 as t E T, t --t 00; (7.4) = {CPTA: (T, A) E G*}, where G* = {(T, A): T;?; to, A E ~,T-llnA E ~ n G}, cp{U) ;2 U for every cP E and if a sequence {(Ti' Ai): i = 1,2, ... } ~ G* is such that Ti --t 00 as i --t 00 and all limit points of the sequence Ti-Iln Ai belong to E, then the sequence of spaces {(CPTiAi{Z))V: i = 1,2, ... } converges in V to the space {Zo)v, where Zo denotes the solution space of the linear equation y' = Ay; (7.5) the space
~n
is represented as the sum
~n
=
~nl
E9 ~n2 (n
= nl + n2),
A = diag{AI' A 2), where Ai' i = 1,2, is a ni x nrmatrix VI < V2:
A=
~
0
o
IA21
Lemma 7.1. Let condition (7.1) hold and 8 > O. Then we have [a, a + 8] x OeD ~ U for some c > o. Proof. The proof consists in referring to Lemma 2.4.2. • Lemma 7.2. Let condition (7.1) hold. Let 8 > 0, c > 0 and [a, a + 8] x QeD ~ U. Then there are numbers CI, 'TJ E (O, c), CI > 'TJ, such that if z E a = inf7r{z), Ilz{a)1I = CI, then SUP7r{z) > a + 8 and z{[a, a + 8D ~ QeD \ [Of/OJ. Proof. Let Z E Zo, a = inf 7r{z) and IIz{a)1I =F O. By the estimates of §IV.5 in [Hal IRe{p, Ap)1 ~ vllpll for some V > 0 for all p E ~n. Here Re x denotes the real part of x. This implies that
zt,
if 8 > J.t > 0 and a + J.t E 7r{z), then Ilz{a)lle- vc5 ~ Ilz{t)11 ~ Ilz{a)lle vc5 {see the argument of §IV.5 in [HaD. Thus the imposed condition holds for CI = ce- 2vc5 and 'TJ = ce- 4vc5 • The lemma is proved. • Lemma 7.3. Let conditions (7.1 - 4) hold. Then IIz{t)11 =F 0, beginning with some t = tl ;?; to. Proof. Take an arbitrary 8 > 0 and select c > 0 according to Lemma 7.1. Next, select Cl > 0 according to Lemma 7.2.
462
CHAPTER 15
For t > to when Ilz{t)11 (7.3) the point
i=
°
put r{t)
= t - a and A{t) = cl1llz{t)II. By
+ In IIz{t)II)(t - a)-I -{lncd{t - a)-l + {In IIz{t)II){t -
r-l{t) In A{t) = {-In Cl =
a)-I,
belongs to the set G, beginning with some t = tl ~ to. Condition (7.4), Lemma 7.2, and Theorem 7.2.2 imply that IIz{t)II beginning with some t = t2 ~ t i . The lemma is proved. Lemma 7.4. Let condition (7.1 - 5) hold,
I Re{x, A1x)1 I Re{y, A 2y)1
~ vdlxll 2 ~ v211yll2
In IIyll -In IIxil
J{p) = {
-00 00
for all for all for for for
i= 0, •
x E ~nl, y E ~n2,
IIxll, IIyll i= 0, IIxil i= 0, IIyll = 0, IIxil = 0, IIyll i=
°
Jorp = {x,y)E ~n\{O}. Then either tlim J{z{t))=-oo or tlim J{z{t))=oo. ..... oo ..... oo Proof. I. By Lemma 7.3 IIz{t)II i= 0, beginning with some t. Assume that the equality lim J{z{t)) = -00 is false. t ..... oo For some mo E ~ the set T{mo) = {t: t E 7r{z), J{z{t)) ~ mol satisfies the condition sup T{ mo) = 00. II. Two cases are possible: A. For some m E (mo, (0) the set T{m) is bounded. B. supT{m) =
00
for all m E (mo, (0).
III. In the case A take 8 = (m - mo) / (V2 - vd and select C according to Lemma 7.1. Next, select Cl according to Lemma 7.2. Let z E Zo, s = inf7r(z) and IIz(s)II i= 0. A direct calculation allows to check that (f{Z{S)))' ~ V2 -VI. Therefore if J{z{s)) ~ mo, then J{z{t)) ~ m for t E 7r{z) n [s + 8, (0) (see also the arguments of the beginning of §X.4 in
[Hal). All limit points of the expression r-l{t)A{t) as t - t 00 belong to E (see the proof of Lemma 7.3). By virtue of the invariance of the function J with respect to the change of variables CPr)" of (7.4) and of our remarks sup T( m) = 00. This contradicts A. Thus the case A is impossible. IV. In the case B we have two possibilities:
c.
sup(~
\ T(m))
= 00 for some m > mo;
Behavior of solutions. D. For all m > mo the set
~
\ T(m) is bounded.
V. In the case C when we take arbitrary mi either: Cl. sup{t: t E 7r(z), J(z(t))
463
= mI,
J(z([t, t
> m and 8 > 0, we obtain
+ 8]))
~ [m,ml]}
= 00,
or:
C2. sup{t: t E 7r(z), J(z(t)) = mI, J(z([t, t + 8])) :3 m} =
00.
In both cases by arguments analogous to III we obtain a contradiction. VI. In the case D limJ(z(t)) = 00. The lemma is proved. •
Lemma 7.5. Let conditions (7.1 - 5) hold. Let
vdlyl12
~ I Re(y,A 2y)1 ~ v211yll2 Jor all y E ~n,
z E Z+, x(t) and y(t) be the components oj the Junction z(t) in ]Rnl and ]Rn2 respectively, IIx(t)1I = o(lIy(t)II). Then limit values oj the expression t-lln IIz(t)1I as t -+ 00 lie in the segment [VI, V2]. Proof. I. By Lemma 7.3 IIz(t)1I :f. 0, begining with some t = to. II. For p = (x, y) E
~n
put
J(p) = J(x, y) =
{~!:"
for for
lIylI:f. 0, lIyll = o.
III. Since IIx(t) II = o(lIy(t)II), lIy(t)1I :f. 0, begining with some t = tl ~ to· IV. Take arbitrary 8 > 0 and select £ > 0 according to Lemma 7.l. Next, select £1 > 0 according to Lemma 7.2. Take arbitrary", > o. V. If Zo E Zo, s, s + 8 E 7r(zo), then vI 8 ~ J(zo(s
+ 8)) - J(zo(s))
~
V28
(see the step III of the proof of Lemma 7.4). Since the expression J(z(s + 8)) - J(z(s)) is invariant with respect to the change of variables CPr>., by (7.4) (VI - ",)8
< J(z(t + 8)) - J(z(t)) < (V2 + ",)8,
beginning with some t = tl
J(z(t»
+ (VI
~
to, or
- ",)8 < J(z(t
+ 8)) < J(z(t» + (V2 + ",)8.
VI. With the notation of IV and V for k = 1,2, ...
J(z(t»)
+ k(VI
- ",)8 < J(z(t + k8)) < f(z(t»)
+ k(V2 + ",)8.
464
CHAPTER 15
VII. By VI for t E [t1
+ k8, t1 + (k + 1)8],
where k = 1,2, ... , we have
f(z(t - k8)) + (t - td(Vl -1/) + (k8 - (t - t 1))(Vl - 1/) = f(z(t - k8)) + k(Vl -1/)8 < f(z(t)) < f(z(t - k8)) + k(V2 + 1/)8 = f(z(t - k8)) + (t - td(V2 + 1/) + (k8 - (t - td)(V2
+ "')
or where
+ 8])) - IV1 -1/18, supf(z([t1, t1 + 8])) + IV2 + 1/18.
m1 = inf f(z([t 1, t1 m2 = VIII. By VII
(t - tdr1(Vl - "')
+ m1r1 < r1ln Ilz(t)1I < (t -
tdr1(V2
+ 1/) + m2rl.
Since Ilx(t)1I = o(lly(t)lI) we have: lIy(t)1I :::; IIz(t)11 :::; 21Iy(t)ll, beginning with some t E 7l"(z). Then
Thus limit values of the expression r-1(t)lnllz(t)11 as t - t 00 lie in the segment [VI - 1/, V2 + 1/]. In view of the arbitrariness of", > 0 this implies what was required. The lemma is proved. • By remarks of the beginning of §X.4 in [Ha] and by the invariance of our condition with respect to a change of basis in the space ]Rn Lemmas 7.4 and 7.5 imply: Theorem 7.1. Let conditions (7.1 - 4) hold, z E Z+. Then the limit lim t- 1 ln IIz(t)1I exists and is equal to the real part of one of eigenvalues of
• the matrix A. Theorem 7.2. Under the hypotheses of Theorem 7.1 let the space ]Rn be represented as the sum ]Rn = ]Rnl EB ]Rn2 EB ]Rn3 (n = n1 + n2 + n3), 1-£1 > 1-£2 > 1-£3, the matrix A have the form
o A=
o
Behavior of solutions.
465
where real parts of eigenvalues of the (n2 x n2)-matrix A2 be equal to J.L2' real parts of eigenvalues of the (nl x nd -matrix Al be not less than J.LI, real parts of eigenvalues of the (n3 x n3)-matrix A3 be not greater than J.L3, z( t) = (ZI (t), Z2 (t), Z3 (t)), and lim t- I In IIz( t) II = J.L2. Then t-+oo
• Example 7.1. Let a non-negative function 1j;(t) be locally integrable on the half interval [a, (0) and
}~
8+6
J
1j;(t)dt = 0
8
for every 6> 0, F(t,y) = {u: Ilull ~ 1j;(t)lIyll}, U = (a, (0) x ]Rn. Let A and Zo be defined as in Theorems 7.1 and 7.2. Let Z denote the solution space of the differential inclusion yl = Ay + F(t, V). The space Z is invariant with respect to the action of the mappings 'Po)., >. > 0, and
{
Xl
= -2x -
yl = _yo
x3t (x 2 _ y2)t 2 + y2
,
Let Zo be the solution space of the system: = -2x, yl = _yo
Xi {
Under the change of variables 'Pr). the first system goes into the system:
= -2x _ x 3 (t + 7) yl = _Yo {X 2 - y2)(t + 7)2 + y2' Xi
{
The convergence of the solution space of the last system to the space Z on the set x 2 - y2 =F 0 follows from classical theorems, see §7.1. In order to prove the convergence on the ray 1 = {(x, x): x > O} consider the part U of the first quadrant lying between the rays x = 2y as
7 --t 00
CHAPTER 15
466
and x = y/2. Consider the set
lnx{t) -lnx{s) t-s
--~~----~~ -
2
~
Cs{U) of all functions z{t) = (x{t), y{t))
for
t> s
and
y'
= -yo
Evidently the set
8. Behavior of a solution space as
Ilyll
---+
00
The situation related to the estimation of the behavior of a solution space as Ilyll ---+ 00 is symmetric with respect to the investigation of stationary points in the 'first approximation' considered in §§6,7. Consider the change of variables .{t,y) = (t - T,Ay). Now we will watch what happens when A ---+ O. Theorem 8.1. Assume that
Z E Ace{lRn) and .{Z) = Z for every A > 0 (i.e., the space Z is conic);
(8.1)
(8.3) the sequence of the spaces 'i (X) converges to the space Z for every sequence {( Ti, Ai): i = 1,2, ... } satisfying the condition Ai ---+ 0; (8.4) there exists a bounded neighborhood G of the point 0 such that if z E Z+, z{inf7r{z)) E [G], thenz{[t,oo)n7r{z)) ~ Gfor some point t E 7r{z).
Then: (8.5) SUP7r{x) = 00 for every function x E X+; (8.6) there exists a compactum K C x{[t, 00)) ~ K for a point t E 7r{x).
]R.n
such that if x E X+ then
Proof. I. By virtue of the boundedness of the set [G] for some a > 0 the set Ga = {au: U E G} contains the compactum [G]. II. Let x E X+ and 7r(x) = [a, b). Prove the boundedness of the function x.
467
Behavior of solutions. Assume the opposite. Then for every k Sk, tk E 7r(x) such that
= 1,2, ... we can fix points
7r(ZZ) = [0, Ck], where Ck = tk - Sk, zZ(O) E [G],
zZ(Cd
ct Gk+o.,
zZ«O, cd) ~ Gk+o. \ [G].
Assertion 1004.1 and the convergence (8.3) imply the existence of a function z E Z+ such that: inf 7r(z) = 0, z(O) E [G] and z(t)
ct G for t > O.
This contradicts (804). Thus our assumption is false. So the function x is bounded. III. Now II and Lemma 6.6.1 imply the fulfilment of (8.5). IV. Prove the existence of a number (3 > 1 such that if x E X+ then x(7r(x) n (t, 00)) ~ G{3 for a point t E 7r(x). Assume the opposite. Then for every k = 1,2, ... there exists a function Xk E X+ such that sup{t: t E 7r(xd, Xk(t) ct Gk(k+o.)} = 00. Let IL(k, t) = inf{(3: (3;:: o!, Xk([t, 00)) n G{3 i= 0} for k = 1,2, ... and t E 7r(xd. By II IL(k, t) < 00. Since IL(k,p) ~ IL(k, q) for p < q, the limit ILk = limt--->oo IL(k, t) exists. By II ILk < 00. The sequence {ILk: k = 1,2, ... } contains a subsequence {ILk: k E A} having a finite or infinite limit IL. Two cases are possible.
A. IL < 00. Consider a subsequence {ILk: k E Ad, where Al For k E Al fix points Sk < tk from 7r(xd such that
=
{k: k E A, k
> IL}.
The existence of such functions x k I[Sk,tlc 1 leads to the same contradiction as in II.
B. IL = 00. The definition of the number ILk implies that for every k = 1,2, ... there exists a point tk E 7r(Xk) such that IL(k, t) ;:: ILk - 1 for t ;:: t k.
CHAPTER 15
468
Z
Let Zk = rptkl-'/:I (Xk l[tk,ooJ· Condition (8.3) and Assertion 10.4.1 imply the existence of a function E Z+, satisfying the conditions:
7r{Z) = [0,00), z{O) E BG and Im{z)
~
IRn \ G.
This contradicts (8.4). The contradiction obtained gives what was required. The theorem is proved. • Example 8.1. Let X be the solution space of the system: {
(8.7)
Xi = -2x + y + 5xe- y2 , yl = -x - 2y - 5xe- y2 ,
Z be the solution space of the system: {
Xi = -2x + y, yl = -x - 2y,
G = {u: u E IRn, lIuli < I}. The change of variables CPT>' transforms the space X into the space X>. of solutions of the system Xl = - 2x + y + 5xe- y, \ -2 2 yl = -x - 2y - 5xe-" y. >.-2
{
2
The convergence X>. - t X may be proved in our usual way, see §§7:3 and 7.5. Thus (8.3) holds. The fulfilment of the hypotheses of Theorem 8.1 is obvious and we obtain from the theorem, that the space X of solutions of the system (8.7) satisfies (8.5) and (8.6).
CHAPTER 16
TWO-DIMENSION AL SYSTEMS
In this chapter we transfer to the framework of our axiomatics several classical results about two-dimensional systems and we continue the topic of the previous chapter in relation to the plane case. As in other analogous situations we obtain here, in particular, an expansion of the corresponding part of the classical theory to equations with discontinuous right hand sides and to differential inclusions. 1. Existence of a stationary point
Let V be an open subset of the plane. Lemma 1.1. Let Z E Ace(V), and K be a minimal (i.e., does not contain a proper subset with the same properties) compact subset of the set V of infinite diameter with respect to the space Z. Then there exists a constant or periodic function z E Z-+ such that K = Im(z). Proof. 1. By Lemma 15.1.1 there exists a function Zo E Z- with Im(zo) ~ K and 7r(zo) = [0,00). By the Corollary of Theorem 15.2.1 diamz n+(zo) = 00. The minimality of the compactum K implies the equality n+(zo) = K. By Theorem 15.2.1 for every point y E K there exists a function z E Z-+ with 7r(z) = (-00,00), z(O) = y and Im(z) ~ K. As in the previous remark the minimality of the compactum K implies the equality n-(zo) = n+(zo) = K. II. If the compactum K consists of one point only, then this point is stationary (see §15.1) and we have what was required. III. Let the compactum K consist of more than one point. Assume that the required periodic function does not exist. Take an arbitrary point Yo E K and a number c > 0 such that K \ [O(Yo, c)] i- 0. Let KI = K n [O(Yo, c)]. By virtue of the minimality of the compactum K diamz KI < 00. IV. If a curve z(t) satisfies the hypotheses ofI and has self-intersections, then the construction of the needed periodic function is easy. So we can require in addition that all such curves have no self-intersections. V. Let Xl = {z: Z E Z, SUP7r(z) = 0, Im(z) ~ K I , z(O) = Yo}, X 2 = {z: Z E Z, inf7r(z) = 0, Im(z) ~ K 1 , z(O) = Yo}.
CHAPTER 16
470
The conditions diamz K1 < 00, Z E Ace(V) and Theorem 3.5.4 imply the compactness of the sets Xl and X 2 . VI. The condition diamz K1 < 00, I and Lemma 6.6.1 imply the nonemptiness of the sets M1 = {z(inf?f(z»: Z E Xd n 8 1 and M2 = {z(SUP?f(z»: Z E X 2} n 8 1 , where 8 1 = O/(Yo,c) \ O(Yo,c) denotes the circle of the radius c with the center at Yo. Theorem 3.5.4, V and Corollary of Theorem 1.7.8 imply the compactness of the sets M1 and M 2. Theorem 1.6.3 implies their closedness in the circle 8 1 . The sets M1 and M2 are disjoint. This follows from IV (or from Lemma 15.1.3). Lemma 2.3.9 implies the existence of disjoint neighborhoods OM1 and 0 M2 of the sets M1 and M 2, respectively. VII. By the Corollary of Theorem 2.6.8 the sets OM1 and OM2 may be represented as the unions of families 81 and 82 of nonempty pairwise disjoint open connected subsets of the circle 8 1 . Evidently elements of the families 81 and 82 are intervals (i.e., arcs without endpoints) of the circle 8 1 (see Theorem 2.6.3 and Example 2.6.5). VIII. Let 8~ = {l: 1 E 81, lnMl i- 0} and 8~ = {l: 1 E 82, lnM2 i- 0}, G 1 = U8~ and G 2 = U8;. Show that there exists a number 8 E (0, c) such that if Z E Z, Im(z) ~ K 1, ?fez) = [a, bj, {z(a), z(b)} ~ 8 1 and Im(z) n O(Yo, 8) i- 0, then z(a) E G 1 and z(b) E G 2. In fact, if we assume the opposite then for every i = 1,2, ... we can fix a function z; E Z such that
= [ai,bi],{Z;(ai),z;(b i )} ~ 8, and either z;(ai) rt. G 1, or z;(bi ) rt. G 2.
Im(z;) ~ K1,?f(Z;) Im(z;)
n O(Yo, 2- c) i- 0 i
By virtue of the autonomy we can put in addition that ai
<
0
<
bi and
z;(O) E O(Yo, 2- i c). By virtue of the equality N = Al U A 2, where Al = {i: i E N, z;(ai) rt. Gd and A2 = {i: i E N, z;(bi ) rt. G 2}, at least one of two sets Al or A2 is infinite. Put for definiteness that the set Al is infinite. By virtue of the conditions Z E Ac(V) and diamz K1 < 00 the sequence {z;: i E AI} has a subsequence {z;: i E A~} converging to a function z* E Z. Evidently Im(z*) ~ K 1, z*(inf ?f(z*» E 8 1 \ G 1, 0 E ?f(z*), z*(O) = Yo. By the definition of M1 we have z* (inf ?f( z*» E MI. Thus our assumption is false, which gives what was required. IX. Fix an arbitrary function Zl E Z+ such that ?f(zd = [0, (0), Zl (0) = Yo and Im(zd ~ K (see I). Since diamz Kl < 00, for some value of the argument t the curve Zl (t) leaves the set O(Yo,c) and then (see I) reaches anew the set O(Yo, Let to = inf{t: t ~ 0, Zl(t) ¢ O(Yo,c)}, b = inf{t: t ~ to, Zl(t) E O/(Yo,
£).
£)}
Two-dimensional systems.
471
and a = sup{t: t E [to,b], ZI(t) E K \ O(Yo,c)}. By the choice of b we have ZI (a) E G 1 • Let 1 be the interval of the family s~ containing the point ZI (a). By the definition of s~ there exists a function Z2 E Z such that Im(z2) <;;; K 1 , Z2(0) = Yo and z2(inf7r(z2)) E 1. Let c = inf{t: t E 7r(Z2), Z2([t, 0]) <;;; O(Yo,c)}. Define the function
z(t) = {ZI(t) Z2(t)
for t E [0,00) for t E [c,O]
on the half interval [c, 00 ). Evidently Z E Z+. Let Q be a point of the set z([O, a]) n 1 that is nearest to the point z(c) (see Lemma 2.3.8). Let q be the corresponding value of the parameter t. Denote by II the part of 1 lying on the circle 8 1 between z(c) and Q. Let 12 be one of two interval of the circle 8 2 of radius %and with center Yo lying between z(b) and z(d), where d = inf{t: t E [c, 0], z(t) E Kl \ O(Yo, %)}. Let P be a point of the set [l2] n z([d, to]) being nearest to the point z(b) (see Lemma 2.3.8), p be the corresponding value of the parameter t.
z
......
Yo
/
'-
I / /'
Figure 16.1
Denote by Uo the region bounded by the closed path Ll consisting of the curves z(t), c:::;; t :::;; p, and z(t), q :::;; t :::;; b, of the interval 11 and of the part 13 of l2 lying on the circle 8 2 between z(b) and P. By Theorem 2.7.2 the closure of the region Uo is homeomorphic to the closed disc, moreover under a corresponding homeomorphism the region Uo goes onto the open disc. Therefore we can joint the points Q and P in the set [Uo] by a curve f(t), t :::;; 1, meeting the boundary of the set Uo at two points Q = f(O) and P = f(l) only. Moreover, the points z(c) and z(b) (see their positions
°: :;
472
CHAPTER 16
on the path L l ) are on different sides of the path L2 constituted by the curves z(t), p ~ t ~ q, and f(t), ~ t ~ 1, (i.e., these points lie on the plane in different components of the complement to L2)' Let U l be the component of the set ]R2 \ L2 containing z(b) and U2 be the component containing z(c) (Figure 16.1). Since the points z(b), z(c) lie in n+(z) (see I), we can then select points SI < tl < S2 < t2 < ... of the interval (b, (0) such that Z(Si) --t z(b) and
°
z(t i ) --t z(c). We have Z(Si) E Ul and z(t i ) E U2, beginning with some i = i o. As the curve z(t) has no self-intersections then either z(t i ) E Uo, beginning with some i = il ~ i o, or for an infinite set A of indices i ~ io the intersection II n Z([Si' til) is nonempty. In each of these cases (for i ~ il in the first and for i E A in the second) we can select points t; E lSi, til such that the sequence {z(t:): i ~ il(i E A)} converges to a point of the set [ld and Z([Si' t:J) ~ Uo U Ul . Let s; = inf{t: t E lSi, ti], z([t, til) ~ [Uon for the same index i. Evidently z([si, t:J) ~ [Uo] and either si = Si, or
z(s:) E O,(Yo, ~). Let a = {z;: i ~ il (i E A)}, where 7f(zi) = [0, ti - s;] and z;(t) = z(si + t). Since diamz Kl < 00 and Z E Ac(V) the sequence a has a subsequence converging to a function z* E Z. We have z*(o) E O,(Yo,~) and z* (sup 7f(z*)) E [ld. This contradicts the conditions z* (sup 7f(z*)) E M2 and II ~ I E S;. Thus our assumption that the compactum K contains more than one point and does not coincide with the set of values of a periodic function • from Z-+ leads to a contradiction. The lemma is proved. Remark 1.1. With the notation of Lemma 1.1 let the set K contain more than one point. Then the set 0, [O,b] E 7f(
°
there exists a function z E Z such that the curve z(t) is a simple loop and the bounded component of the set ]R2 \ Im(z) (see Jordan theorem 2.7.1) coincides with U. Lemma 1.2. Let Z E Ace(V) and Uo E P(Z, V). Then either the set
[UoJ contains a stationary point of the space Z, or there exists a set U E
Two-dimensional systems.
473
P(Z, V) such that U ~ Uo and U does not contain elements of P(Z, V) as proper subsets. Proof. The proof is analogous to the proof of Lemma 15.1.4. Fix a countable base f3 = {Bi : i = 1,2, ... } of the set [Uo]. Sequentially construct elements Uo :2 U1 :2 U2 :2 U3 :2 ... of the set P(Z, V). Let the sets Uo, ... , Ui be fixed. Two cases are possible. A. There is an element U' of the set P(Z, V) such that U' ~ Ui \ B i . Put UiH = U'. B. The set Ui \ Bi does not contain elements of P(Z, V). Put Ui+l = Ui . The set Ko = n{[Ui ] : i = 1,2, ... } is compact and lies in [Uo]. By the Corollary of Lemma 15.1.2 diamz Ko = 00. By Lemma 15.1.4 there exists minimal a compactum K ~ Ko of the infinite diameter with respect to Z. If the set K is one point then we have what was required. Let the set K contains more than one point. By Lemma 1.1 and Remark 1.1 there exists a function z E Z such that Im(z) = K and z is a simple loop. By Jordan theorem 2.7.1 the simple closed arc z bounds a region U. Show that U ~ lUi] for every i = 0,1,2, .... Assume the opposite. Then for some i = 0,1,2, ... the set U \ lUi] is nonempty. Since the sets lUi] and [U] are compact they do not cover the entire plane. Therefore the set (]R2 \ lUi]) \ U is nonempty. By virtue of the connectedness of the set ]R2 \ lUi] and the nonemptiness of the sets (]R2 \ lUi]) n U = U \ lUi] and (]R2 \ lUi]) \ U the set U \ lUi] may not be closed in ]R2 \ lUi]. Therefore its boundary K \ lUi] is nonempty. This contradicts the choice of the set K. This gives what was required. Since the set U is open, U ~ ([Ui ]) = Ui for every i = 0,1,2, .... This means, that U ~ n{Ui : i = 0,1,2, ... }. It remains to show that the set U minimal, i.e., it does not contain elements of the set P(Z, V) as proper subsets. Assume the opposite, i.e., that there exists a set U* E P(Z, V) being a proper subset of U. Let z* E Z be a simple loop bounding U*. The arc Im(z*) is the common boundary of the regions U* and ]R2 \ [U*] (see Jordan theorem 2.7.1). This implies, that the set U \ [U*] is nonempty. Let Bi be a (nonempty) element of the base f3 lying in U \ [U*]. Then the set Ui \ Bi contains U*. Therefore UiH ~ Ui \ B i . Since the set U lies in the set UiH , U ~ ]R2 \ B i . This contradicts the choice of Bi (Bi ~ U). The obtained contradiction proves that the set U does not contain elements of P(Z, V) as proper subsets. The lemma is proved. • Lemma 1.2 is not complete: in fact always the first of the mentioned two possibilities take place, namely Theorem 1.1. Let Z E Ace(V) and Vo E P(Z, V). Then the set [Vo] has a stationary point of the space Z.
474
CHAPTER 16
Proof. 1. By the definition of P(Z, V) there exists a function Zo E Z such that the curve zo(t) is a simple loop bounding the region Vo. By Lemma 1.2 it is sufficient to consider the case when diamz M < 00 for every compact subset M of the set [Vo], which does not contain the entire boundary I = Im(zo) = a[Vo]. Theorem 2.7.2 and remarks of §10.1 imply that we can assume in addition that I is the unit circle with center at the origin of coordinates and Vo is the open disc bounded by this circle. The proof will consist in the consideration of two cases. In the first we shall assume that for a function Zl E Z the intersections Im(zd n I and Im(zd n Vo are nonempty. In the second we shall assume that such a function does not exists. II. Consider the first case. Evidently we can assume in addition that the function Zl takes values in the circle I only at one of the endpoints of the segment 7r(zd. Assume that 7r(zd = [al' bd and zl(al) E I (the case Zl (bd E I may be considered in an analogous way or we can obtain the corresponding result from the previous case after the change t --t (-t)).
Figure 16.2
Extend the function Zl on the right to a function Z2 E Z+ (see Lemma 6.6.2). If Z2(t) E I for some t E 7r(Z2) \ {ad, then we can use the functions Zo and z21[at ,t] and the condition of the autonomy of the space Z to construct a periodic function ( E Z-+ such that values of ( lie in [Vo] and do not cover entirely the set l. This contradicts the assumptions of 1. Thus the curve Z2(t), t E 7r(Z2) \ {ad, does not meet the circle l. The connectedness of the set 7r( Z2) \ {ad implies that Z2 (7r( Z2) \ {ad) ~ Vo. The compactness of the set Va U l, the condition z E Ac(V), and Lemma 6.6.1 imply that
Two-dimensional systems.
475
sup 7r(Z2) = 00 and 7r(Z2) = [aI, 00). By analogous arguments and by Lemma 15.1.3 the curve Z2(t) has no self-intersections. Fix a number c: E (0,1) such that Ilz2(bl) - z2(adll > 2c:. The set O(z2(ad,2c:) cuts on the circle I the arc 120 with 0 and (3 as endpoints. The arc 120 contains the point z2(al). The intersection 120 n Im(z2) consists of one point z2(ad. By the Corollary of Theorem 15.2.1 diamz n+(Z2) = 00. Therefore by the assumptions of I 120 ~ n+(Z2). Denote by If.> and IfJ the intervals lying in the arcs 120 and having 0 and z2(ad, {3 and z2(ad, respectively, as endpoints (Figure 16.2). Let s~ = inf{t: t E 7r(Z2), Ilz2(t)-z2(adll = 2c:}. Let Sf.> and SfJ be the intervals of circle of the radius 2c: with the center at the point z2(al), which lie in the disc Yo and which have 0 and Z2(S~), (3 and Z2(S~), respectively, as endpoints. The set O(z2(al), c:) cuts on the circle I the interval 10 with the endpoints 01 (on the side of the point 0) and {3l (on the side of the point (3). Let If be the part of the arc 10 with the endpoints (3l and z2(ad. Construct sequentially points tt < t; < st < s;, i = 1,2, ... , of the set 7r(Z2)· Let the point SLI E [SfJ], i = 1,2, ... , be chosen and the part of the arc SfJ between the points z2(sLl) and (3 have no points ofthe curve Z2(t), al ~ t ~ sL 1. Denote by B the region bounded by the arc IfJ, by the curve Z2 (t), al ~ t ~ SLll and by the part of the arc SfJ lying between the points z2(sLl) and (3. Since the arc If lies in the set n+(Z2), the curve Z2(t), t > SLll must enter in the intersection A of the set O(z2(al), c:) and the region B. Let t; = inf{t: t> SLl' Z2(t) E A}. Let tt be the value of the parameter t E [SLl' t;] corresponding to the point of the intersection of the curve z(t), SLI ~ t ~ t;, and of the arc [SfJ], being nearest to (3. Denote by Bl the region bounded by the arc IfJ, by the curve Z2(t), al ~ t ~ tt, and by the part of the arc SfJ, lying between the points Z2 (tt ) and (3. Since tf.> ~ n+ (Z2), the curve Z2 (t), t > t~, must leave the region B l . Let s~ = inf{t: t;;:::: t~, Z2(t) ~ Bd. Since the curve z(t) has no selfintersections, Z2 (sD E SfJ. The part of the arc SfJ between the points Z2 (si) and (3 has no points of the curve Z2(t), al ~ t ~ sr Denote by st the value of the parameter t E [t~, s;], corresponding to the point of the intersection of the curve Z2(t), t; ~ t ~ s;, and of the circle of the radius c: with the center at the point Z2 (al) being nearest to (3l. Evidently under this construction the points Z2(S~), z2(tD, Z2(Si), Z2(t~), Z2 (s~), . .. move monotonically in the arc SfJ in the direction from the point z(s~) to the point (3 and the points Z2(t~), z2(sD, z2(tD move monotonically in the circle of the radius c: with the center at the point z2(ad from the point 01 to the point {31.
476
CHAPTER 16
Let
and
°
Since diamz O,{z2{ar), 2c) n [vol < 00, for some M > we have Is~ - s;1 :::; M and It~ M for every i = 1,2, .... For i = 1,2, ... on the segments [0, s; - s~l and [0, respectively, define the functions z; and z;* by the formulae z;{t) = Z2{S~ + t) and
m: :;
z;*{t) = z2{ti
t; - tn
+ t).
Since Gr{z;), Gr{zt) ~ [0, Ml x [Vol, i = 1,2, ... , and Z E Ac{V), the sequences {z;: i = 1,2, ... } and {z;*: i = 1,2, ... } have subsequences converging, respectively, to functions z*, z** E Z. The initial point of the curve z*{t) (the point 12) coincides with the end of the curve z**{t), and the end of the curve z*{t) (the point II) coincides with the initial point of the curve z**{t). In view of Lemma 15.1.3 and the membership Z E A{V), this contradicts the finiteness of the diameter of the set O,{z2{ar), 2c) n [Vol with respect to the space Z. Thus the first case is impossible. III. Consider the second case: If z E Z, Im{z) ~ [Vol and Im{z) then Im{z) ~ l.
n 1 =1= 0,
The curve 1 may be run by functions from Z in one direction only because in the opposite case it is not a minimal set of the infinite diameter with respect to the space Z. For definiteness put this direction counterclockwise (the other case may be considered in an analogous way or we can obtain the corresponding result from the following consideration with the help of the change t = -s). IV. Let J 1 , J 2 and J 3 be successive (for the counterclockwise moving) E 7r{Zi) for i = 1,2, ... , coordinate rays, {Zi : i = 1,2, ... } ~ Z-+, Zi{O) E J 2 , Ilzi{O)11 < 1 and Ilzi{O)11 ~ 1. Since the curve Zi{t), i = 1,2, ... , does not leave the disc Vo Lemma 6.6.1 implies that 7r{Zi) = (-00,00). Let Q 2 J 2 be the open half plane with the boundary J 1 U J 3 , C E (0,1), Ve be the open disc with the center at the origin of coordinates and the radius 1 - c, Qg = Q n (Vo \ [Ve]). Show that, beginning with some i, we can fix segments [Si, til ~ 7r(Zi), such that Si :::; ti, Zi([Si, ti]) ~ [Qg], Zi([Si, ti]) n [Vel = 0, Zi(Si) E J 1, Zi(ti) E J3, Zi([Si, 0]) n J 3 = 0 and
°
°: :;
Two-dimensional systems.
477
n J l = 0 (Figure 16.3). Notice first that Zi(O) E Qo beginning with some i = i o. Therefore by virtue of the finiteness of the diameter of the set [Qo] with respect to the space Z (see I) we can select points Si < 0 < ti such that Zi([Si, ti]) ~ [Qo] and Zi([O, tiD
Zi(Si), Zi(t i ) E 8Qo· Show that the intersection n [Vo] is empty, beginning with some i = i l ~ i o. Assume the opposite. By virtue of Figure 16.3 finiteness of the diameter of the set [Qo] with respect to the space Z numbers ISil, ti, i = i o, io + 1, ... , are bounded in the totality by a number M > 0 and the graphs of the functions zilrsi,t;j lie in the compact urn [-M, M] x [Qo]. By virtue of the condition Z E Ac(V) we can pass to a subsequence of the sequence {zilrsi,t;j : i = i o, io + 1, ... , Zi([Si, td) n [Vo] -=I- 0} converging to a function Z E Z. For it z(O) Eland Im(z) n [Vo] -=I- 0, that contradicts the assumptions ofIlI. Show that Zi([Si, 0]) n J 3 = 0, beginning with some i = i2 ~ il. Assume the opposite. As in the previous paragraph the sequence
Zi([Si, tiD
has a subsequence converging to a function Z E Z. For it z(O) Eland Im(z) n J 3 -=I- 0. Two cases are possible. A. Im( z) n Vo -=I- 0. This contradicts the condition in the beginning of III. B. Im(z) ~ l. This contradicts remarks of III about the direction of the moving in the circle l. Thus zi([si,OD n (J3 u [Vo] u 1) = 0. Since Zi(Si) E 8Qo we have the unique possibility Zi(Si) E J l . Likewise Zi([O, tiD n J l = 0 and Zi(t i ) E J 3 beginning with some i = i3 ~ i 2. V. Keep the notations of IV. By virtue of the condition of the autonomy of the space Z the result of IV may be restated as follows: there exists a number 8 E (c, 1) such that if Z E Z-+, a E 7r( z) and z( a) E Q 6 n J 2 , then we can select points S < a < t of the set 7r( z) such that z([s, tJ) ~ [Qo]' z([s, tJ) n [Vo] = 0, z(s) E J l , z(t) E J 3 , z([s, aJ) n J 3 = 0 and z([a, tJ) n J l = 0 (in the opposite case we construct in an obvious way a sequence of functions, properties of which contradicts IV).
CHAPTER 16
478
VI. Keep the notation of IV and V. Let J 4 be the fourth coordinate ray. For E: = ~ find 8 = 81 according to V. Next, for E: = 81 and for the rays J4 , J 1 , J 2 (as the rays J 1 , J 2 , J 3 , which occur in V) find 8 = 82 according to V. For E: = 82 and the rays J 3 , J4 , J 1 find 8 = 83 . For E: = 83 and the rays J 2 , J 3 , J4 find 8 = 84 • ForE: = 84 and the rays J 1 , J 2 , J3 find 8 = 85 (Figure 16.4). Fix an arbitrary function z E Z-+ with 0 E 7r(z) and z(O) E Q0 5 n J 2. According to V and to our choice of 85 we have
z([t o, 0]) to < o.
~
[Q04]' z([to, O])
n ([V6 4] u J3 U l)
= 0 and z(t o) E J 1 for some
According to V and to our choice of the numbers 85 ,84 ,83 ,82 and 81 the curve z(t), t ~ 0, makes one rotation and one fourth and for some t = t4 > 0 (the first time after the rotation and one fourth) meets the ray J3 .
Figure 16.4
In the first rotation the curve may meet the ray J 3 several times. Let t2 be the value of the argument t E [0, t 4) corresponding to a point z(t) of the intersection z([O, t 4)) n J 3 being nearest to z(t 4) (or one of nearest points if we have two such points). This value t2 exists by virtue of the compactness
479
Two-dimensional systems.
of the set {t: t E [0, t 4J, Z( t) E J 3} and of the isolatedness of the point t4 in the set. For t < 0 and on the segment 0 ~ t ~ t4 the curve z(t) may meet the ray J 1 several times. Let tr E [to, OJ and t3 E [0, t4J be values of the argument t corresponding to points of the intersection of the curve and of the ray J 1 such that the interval with the endpoints z(t 1 ) and Z(t3) does not contain points of the curve
z(t), to
~
t
~
t4·
Compose two paths. The first path Ll consists of the curve z(t), tl ~ t ~ t 3, and of the segment connecting the
Figure 16.5
point z(tr) and Z(t3). The second. path L2 consists of the curve z(t), t2 ~ t ~ t 4, and of the segment connecting the points z(t 2) and z(t 4) (Figure 16.5). Show that the curve z(t), t ~ t 4, does not meet the path L 1 . Assume the opposite. Since the curve z(t) has no self-intersections then the curve z(t), t ~ t 4, meets the interval II of the ray J 1 with the endpoints z(tr} and Z(t3). Let 82 be the smallest value of the argument t E [t4' (0) corresponding to such an intersection. By the definition of 82 the curve z(t), t4 ~ t < 82, does not meet the path L 1 • Therefore it lies entirely in the component U of the set ~2 \ Ll (see Jordan theorem 2.7.1) containing the point z(t 4). Let the path L3 be composed by the curves z(t), tl ~ t ~ t 2, and z(t), t3 ~ t ~ t 4, the interval II and the interval 13 of the ray J3 with the endpoints z(t 2) and z(t 4). The bounded component Uo of the complement in the plane to the path L3 has no common points with the path L 1 . By Corollary 1 of Lemma 2.6.1 it lies entirely in one of the components of the set ~2 \ L 1 • Since z(t 4 ) E auo, Uo ~ U. The interval II splits rather small neighborhoods of its points in two parts. One of the parts lies in the component U* = ~2\[UJ of the set ~2\Ll' and the other one lies in Uo ~ U. Therefore the curve z(t), t4 ~ t ~ 82, may approach to the point Z(82) only from the side of the region Uo. Let 81 = inf{t: t E [t 4,82], Z([t,82]) ~ Uo}. Thus Z(8r} E [UoJ \ Uo. The boundary of the bounded region Uo lies in the set [Qtl, therefore Uo ~ Qt· Both the curve z(t), tl ~ t ~ t 2, and the curve z(t), t3 ~ t ~ t 4, have points of the interval 12 = J 2 n Q5,. This implies that the intersection Uo n J 2 contains the interval 14 of the
480
CHAPTER 16
ray J 2 , one of the endpoints of which lies in the curve z(t), tl ::::; t ::::; t 2 , and the second endpoint lies in the curve z(t), t3 ::::; t ::::; t 4 . Since the curve z(t) has no self-intersections and SI < S2, z(sd E 13, Theorem 2.7.1 easily implies that the interval 14 splits the region Uo in two parts. One of the parts is adjacent to the interval II, The other one is adjacent to the interval 13 , Corollary 1 of Lemma 2.6.1 implies now the nonemptiness of the intersection Z«SI' s2))nI4. Thus Z«SI' S2))nQ62nJ2 i- 0. Now properties of the function ZI[Sl,S2] (the position of the points Z(SI) and Z(S2)) contradicts the choice of 8 = 82 in V. Likewise the curve z(t), t ::::; t I , does not meet the path L 2 • The circle I lies in the unbounded components of the sets ~2 \ LI and ~2 \ L 2 • Two cases are possible. A. The component U ofthe point z(t 4 ) in the set ~2 \ Ll is bounded (see Figure 16.6). By above remarks Z([t4' 00)) lies in U. Therefore n+ (z) ~ [U] and n+(z) n I = 0. By the Corollary of Theorem 15.2.1 this contradicts the Figure 16.6 assumption of I. B. The component U of the point z(t 4 ) in the set ~2 \ Ll is unbounded (Figure 16.6). In this case the union U** of the bounded component of the set ~2 \ Ll , of the region Uo, of the arc Z([tl' t 2]) and of the interval II is bounded. By Jordan theorem 2.7.1 and from the inclusion Uo ~ U the boundary of the set U** coincides with the path L 2 • Therefore [U**] n I = 0. According to above remarks z«-oo,td) ~ U**. Therefore n-(z) ~ [U**] and n-(z) n I = 0. By Corollary of Theorem 15.2.1 this contradicts the assumption of I. The obtained contradiction completes the consideration of the second • case. The theorem is proved. 2. Indices
Let V be an open subset of the plane and Z E Acek (V). Let x E V, c > 0, K = O/(x, c) ~ V, diamz K < 00 and S* be the circle bounding the disc K.
Let M = {z : z E Z, 0 E 7r(z), z(O) = x, Im(z) ~ K, z{inf7r{z)),z{sUP7r(z)) E S*}. The conditions Z E Ace(V), diamzK < 00 and Lemmas 6.6.2 and 6.6.1 imply the nonemptiness of the set M. For
481
Two-dimensional systems.
Z E Z denote a(z) = z(inf7r(z)) and f3(z) = z(SUP7r(z)). Lemma 15.1.3 implies that a(M) n f3(M) = 0. Fix an arbitrary function Zo E M. Denote by rl (respectively, by r2) the set of all segments of the circle S* with endpoints from the set a(M) (respectively, from the set f3(M)) which contains the point a(zo) (respectively, f3(zo)) and which does not contain the point f3(zo)) (respectively, a(zo)). For i = 1,2 put Pi = Uri' The conditions z E Ace(V), diamz K < 00 and Theorem 3.5.4 easily imply the closed ness of the sets PI and P2' Theorem 2.6.1 implies their connectedness. Thus the sets PI and P2 are segments of the circle S*. Evidently Pi E ri (where i = 1,2). Our definition immediately implies that a(M) ~ PI and f3(M) ~ P2' Lemma 2.1. With the above notation PI np2 = 0. Proof. Assume the opposite. Then the set PI n f3(M) is nonempty. Let Xl be an arbitrary point of this set. Fix three functions Zl, Z2, Z3 E Z such that a) Im(zi) ~ K for i = 1,2,3; b) 7r(zd = [0, b] and Zl (b) = Xl; c) 7r(zd = [ai'O] for i = 2,3 and z2(a2), z3(a3) are different endpoints of the segment PI; d) Zi(O) = X for i = 1,2,3. If z* E Z, 7r(z*) = [0, c], Im(z*) ~ K and z*(O) = x, then
Im(z*)
n (Im(z2) U Im(z3)) = {x}.
This follows easily from Lemma 15.3.1 and from the finiteness of the diameter of the compactum K with respect to the space Z. For i = 2,3 let li denote the ray with initial point zi(ai) and direction vector zi(ai) - x. Let Kl be an arbitrary disc with center X and radius ~ c:. Jordan's theorem 2.7.1 easily implies that the complement to the set Im(z2)Ulm(z3)U1 2U1 3 in the disc Kl may be represented as the union of two its components G l and G2 • The set Zl «0, b]) lies in one of this component. The set zo«O,SUP7r(zo)]) lies in the another component. Thus for a small c > 0 the set Ac = {z(c): Z E Z, 7r(z) = [0, c], z(O) = x} lies in G l U G 2 and meets both G l and G2 • Lemma 2.6.1 implies the disconnectedness of the set Ac. The last fact contradicts the condition Z E Ak(V), Therefore our assumption is false and the lemma • is proved. Corollary. The definition of PI and P2 does not depend on the choice of Zoo • Denote
S(x,c:)
= {y:
yE
]R2,
p(x,y)
= c:},
P(x,c:)
= {(u -
x)c:- l
:
U
E P2}
482
CHAPTER 16
and Vp = {(X*, C*): X* E V, c*
> 0, diamz Of(x*, c*) < oo}.
The previous reasoning allows us to consider P as a multi-valued mapping from Vp to the unit circle S. Its values are proper connected compact subsets of the circle S. Lemma 2.2. The mapping P : Vp --t S is upper semicontinuous. Proof. The proof use Theorem 2.5.1. I. Take arbitrary sequences"., = {(Xb Ck): k = 1,2, ... } ~ Vp and Yk E P(Xk' cd, k = 1,2, .... Let the sequence"., converge to a point (xo, co) E Vp. For k = 1,2, ... let uk and uk* denote different endpoints of the segment P(Xk' Ck). Fix functions Zk' zt E Z satisfying the conditions:
o E 7T(Z;) n 7T(Z;*),
z;(O)
= z;*(O) = Xb
Im(z;) U Im(z;*) ~ Of(Xk, cd, z;(inf 7T(Z;)) E S(Xk' cd, Z~(SUp7T(Z~)) = Xk + CkU:, Zt(SUp7T(Z~*)) = Xk + CkU:*. II. Since diamz 0f(xo, co) <
00,
by Lemma 15.1.2
diamz Of(xo, Co
+ 8) < 00
for some 8 > O. The convergence Xk --t Xo and Ck --t Co implies easily that Of(Xb Ck) ~ O(xo, Co + 8), beginning with some k = ko. Therefore all elements of the sequences {Zk: k = ko, ko + 1, ... } and {zt k = ko, ko + 1, ... } belong to the compactum ZM, where
M = Of(xo, Co
+ 8)
x [- diamz 0f(xo, Xo
+ 8), diamz Of(xo, Co + 8)]
Recall that the space Z satisfies condition (c). Pass to subsequences {zk: k E Ai} and {Zk*: k E Ai} converging, respectively, to functions z;,z;* E ZM. The points Ck = Xk + CkYk, k E Al lie in the compactum 0f(xo, Co + 8). Therefore the sequence {Ck: k E Ai} has a subsequence {Ck: k E A 2 } converging to a point C E 0f(xo, Co + 8). III. By Theorems 3.5.1 and 3.5.4 the sequences {zk(inf7T(zk)) : k E A 2 }, {zk(SUp7T(Zk)) : k E A 2 } and {Z;*(SUp7T(Zk*)): k E A 2 } converge, respectively, to the points z;(inf7T(z;)), Z;(SUp7T(Z;)) and z;*(SUp7T(Z;*)). Denote one arbitrary of them or the point c by t. For k E A2 denote by tk one arbitrary of the points zk(inf7T(zk))' Zk(SUp7T(Zk))' zk*(SUp7T(Zk*)) or Ck, respectively. By Lemma 2.3.1 Ip(Xk' t k) - p(xo, t)1 ~ Ip(Xk' t k) - P(Xk' t)1 ~ P(tk' t) + P(Xk' xo).
+ Ip(Xk' t) -
p(xo, t)1
Two-dimensional systems.
483
Therefore p(t, xo) = limk-+oo P(Xk' td = limk-+oo Ck = co. Thus the point t lies in the circle S(xo, co). IV. The points Zk (sup 7r(Zk)) and zA;*(sup 7r(Zk*)) split the circle S(Xk' cd into two arcs. The points Ck and zk(inf7r(zk)) lie in different of these arcs. Therefore the rectilinear segments h with the endpoints Zk (sup 7r (Zk))' Zk*(SUP7r(Zk*)) and J k with the endpoints zk(inf7r(zk))' Ck intersect. Evidently this property is kept under passage the limit. Therefore the segment I with the endpoints z~(SUP7r(z~)), z~*(SUP7r(z~*)) and the segment J with the endpoints z~(inf7r(z~)), C intersect too. The points z~(SUP7r(z~)) and z~*(SUP7r(z~*)) split the circle S(xo,co) in two arcs. The preceding remark implies that the points z~(inf7r(z~)) and C lie in different of these arcs. Since Im(z~) U Im(z~*) ~
lim top sup Im(zZ) U lim top supIm(z~*) k-+oo k-+oo ~ limtopsupOf(xk,ck) ~ °f(xo,co), k-+oo
the segment 1 of the circle S(xo, co) with endpoints z~(sup 7r(z~)) and z~*(SUP7r(z~*)), which does not contain the point z~(inf7r(z~)), satisfies the condition {(u - XO)cOl : u E l} ~ P(xo, co). By the remarks of the previous paragraph the segment 1 contains the point c. Therefore the point (c - xo)co 1
= k-+oo lim (Ck
- xdc;;l
= k-+oo lim Yk
.
belongs to the set P(xo, co). Referring to Theorem 2.5.1 completes the ~~
The set of stationary points of the space Z is denoted by Stat(Z). A mapping f : X --t V of an arbitrary set X in the set V is called free of stationary points (of the space Z), if f(X) ~ V \ Stat(Z). Let now X be a segment [a, bj of the real line or the unit circle S. Let f : X --t V be a loop free of stationary points of the space Z. Evidently the set of stationary points is closed (in V). Therefore the set V \ Stat(Z) is open (in the plane). If V = ~2 and Stat(Z) = 0, then put co = 00. In the opposite case put co = p(f(X), (~2 \ V) u Stat(Z)) (by Lemma 2.3.8 co > 0). Lemma 2.3. diamz 0f(f(x), c) < 00 for every x E X and c E (0, co). Proof. Assume the opposite. By Lemma 15.1.4 there exists a minimal compactum Ko ~ Of(J(x), c) of the infinite diameter with respect to the space Z. By Lemma 1.1, by Remark 1.1 and by Theorem 1.1 the set Ko contains a stationary point of the space Z. This contradicts the definition of co. The lemma is proved. • Let c E (0, co). By the previous reasoning the formula ge(x) = P(f(x), c) defines an e-Ioop or a mapping from e(S, S) (see §2.8). Denote the degree of this mapping by j(Z, f, c).
484
CHAPTER 16
Let 0 < el < e2 < co· The formula G(x, c) = P(f(x), c) defines an e-homotopy G : X x [el' C2] - t S connecting the mappings gel and g£2. By Theorem 2.8.1 j(Z, f, cd = j(Z, f, C2). Denote the common value of j(Z, f, c), e E (0, co), by j(Z, I). It is called the index of the space Z with respect to f. Let F : S x [c, d] - t V be an homotopy connecting the mappings II and 12 and being free of stationary points of the space Z. Let e* = 00 if V = ]R2 and Stat(Z) = 0, and e* = p(F(S x [c, d]), (]R2 \ V) U Stat(Z» in the opposite case. The formula G(x, t) = P(F(x, t), c) (for every e E (0, e*» defines an e-homotopy connecting the mappings gl£(X) = P(F(x, c), c) and g2£ = P(F(x, d), c). By Theorem 2.8.1 the degree of the mappings gl£ and g2£ coincide. So we have proved:
Theorem 2.1. If (continuous) mappings fl' f2 : S - t V may be connected by an homotopy free of stationary points of the space Z, then j(Z, fd = j(Z, h). • If f is a constant mapping, then, of course, j(Z, f) = O. If a simple closed path S - t V bounds a region W, then Theorem 2.7.2 implies easily the existence of an homotopy connecting in [W] the mapping rand a constant mapping. Therefore Theorem 2.1 implies: Corollary. If a simple closed path S - t V (free of stationary points of the space Z) bounds the region W ~ V and j (Z, r) f:. 0 then the region W has at least one stationary point of the space Z. • Let a E ]R2. Let f be a loop with values in ]R2 \ {a}. The degree of the mapping f(t) - a
r :
r :
g(t) = Ilf(t) -
all
is called the index of the point a with respect to f. Let 0 < fJ < 80 < 00. Let the index of the point a with respect to the continuous mapping f : S - t O( a, 80 ) \ {a} is equal to 1. By Theorem 2.8.2 the above mapping 9 is homotopic to the identity mapping. Let G : S x [1,2] - t S be a corresponding homotopy (G(8,1) = g(8) and G(8, 2) = s). The formula
F(8, t) = { f(8)(1 - t) + ( a + a+G(s,t)8
(f(8) - a)8)
Ilf(8) _ all
t
for s E S, 0 ~ t ~ 1, for
8
E S, 1 ~ t ~ 2
defines an homotopy F : S x [0, 2] - t O( a, 80 ) \ {a} connecting the mappings and 10(8) = Xo + 8s. Recall that we do not impose on 8 E (0,80 ) any additional conditions. In addition let O(a, 60 ) \ {a} ~ V \ Stat(Z). By remarks of the previous paragraph and by Theorem 2.8.1 the index of the space Z with respect to I
f
Two-dimensional systems.
485
does not depend on a concrete choice of f and 6 (we assume the fulfilment of all mentioned conditions). This common value j(Z, f) is called the index of the point a with respect to the space Z and is denoted by jz(a). Theorem 2.2. Let the region W ~ V be bounded by a simple closed path f : 8 -+ V. Let the path go around W counterclockwise and do not contain stationary points of the space Z being different from aI, ... ,ak (E W). Then j(Z, f) = iz(al) + ... + jZ(ak). Proof. The proof follows an induction on k. For k = 1 our assertion follows easily from Theorems 2.7.2 and 2.1. Let the assertion is true for k = i. Prove it for k = i + 1. Select a point from al,···, ai+l being
~l
y Figure 16.7
nearest to the arc f(8), i.e., if ajo is such a point, then p(ajo' f(8)) = min{p(aj, f(8)): j = 1, ... ,i + I}. Let y be a point of the arc f(8) being nearest to ajo (such a point exists by Lemma 2.3.8). Assume that 6 > 0 is such that the 26-neighborhood of the point ajo does not contain the points al, ... ,ajo-l, ajo+l, ... ,ai+l and does not meet the arc f(8). Let I be the segment connecting the points ajo and y. Let Ll be a closed path which runs first over the path f from y to y, next the part of the segment I to its intersection with the circle 8(ajo' 6), next the circle 8(ajo' 6) in the negative direction (i.e., so that the index of the point ajo with respect to the loop so run is equal to -1, in our account this takes place under the clockwise motion) and the part of the segment I to the point y (Figure 16.7). Let L2 be a closed path, which starts at the point y and runs first the part of the segment I to its intersection with the circle 8 (ajo , 6), which runs next the circle 8 (ajo' 6) in the positive direction and returns by I to the initial point. Let L be a closed path, which runs first 11 and next 1 2 . The inductive hypothesis easily implies that
486
CHAPTER 16
The definition of the index of a point implies that j(Z, L 2) = jz(ajo). Lemma 2.8.3 implies the equality j(Z, L) = j(Z, Ld + j(Z, L2). We leave to the reader as an exercise the construction of an homotopy (free of stationary points of the space Z) connecting f and the mapping of the circle corresponding L. By Theorem 2.1 j(Z, f) = j(Z, L). Hence
j(Z, f) = Jz(ad
+ ... + jZ(ai+1).
The theorem is proved. • Remark 2.1. The technique of this section allows us to prove Theorem 1.1, following the plan of the proof of an analogous assertion of the classical theory of ordinary differential equation; however, in this way we must assume in addition that the space Z satisfies the Kneser condition. 3. Calculation of index by the right hand side In the previous section we did not ask ourselves in what measure in reality we are able to calculate an index in a particular situation. If we follow the way of §2 we must give first a sufficient description of the space Z. Next we must give a description of the mapping P, which may be sufficient for the calculation of the index. The situation becomes simpler when we deal with the space Z of solutions of an inclusion y' E F(y). In addition to the notation of §2 let the function F E Q*(lR x V) (see §6.4) do not depend on the first argument. A point y E V is a stationary point of the space D(F) if and only if E F(y). In the opposite case by the definition of Q*(lR x V) (see §6.4) there exists a number 'T/ > 0 that such
o
(3.1)
The set A = cc(F(07jY)) is convex and closed. By Theorem 2.4.2 (under the assumption of the nonemptiness of the set F( 07jY)) there exist a linear functional 0 : lR2 -+ lR and real numbers p < q such that O(A) ~ (-00, p] and 0(0) E [q,oo). Since 0(0) = 0, p < q:::;; O. The set {u: u E lR 2 , O(u) :::;; p} is a closed half-plane, which does not contain the origin of coordinates. Thus (3.2) the set F(07jY) lies in a closed half plane which does not contain the origin of coordinates. Assume in addition that the space D(F) satisfies conditions (c) and (e). Since the solution space of the inclusion with the identically empty right hand side does not contain functions with nontrivial domains, condition (e) (for the space D(F» implies, that F(W) f. 0 for every nonempty open set W ~ V. This makes the previous remark valued.
Two-dimensional systems.
487
Theorem 3.1. Under the above assumptions the space D{F) satisfies the K neser condition. Proof. Let Y E V and to E ~. Our aim is to show that there exists a number 8 > 0 such that for every point s E (to - 8, to + 8) and for s = {z: Z E D{F), {inf7r{z),sUP7r{z)} = {to,s}, z{t o) = y}, the set As = {z{s): Z E s} is connected. Since the space D{F) is autonomous, for simplicity of notation we can assume in addition that to = O. For definiteness let s > O. Consider two cases. 1. Let 0 E F{y). If Z E s, 0 < a < s, then the function
Z,.{t) =
{Y( t - a ) Z
for 0 ::; t ::; a, for a ::; t ::; s,
belongs to the set s too. Since {za(s): 0::; a ::; s} = Im(z), in view of the arbitrariness of the function Z E s we have As = U{Im(z): Z E s}. Now the connectedness of the set As follows from Theorem 2.6.1. II. Let 0 rf. F{y). Find 'fJ > 0 satisfying (3.2). Let A be a corresponding closed half-plane. Let the line 1 bound A and p* be the distance from the origin of coordinates to 1 and A. By Theorem 6.7.3 there are numbers f-£ > v > 0 such that f-£ < 'fJ and if Z E (D{F»-+, 0 E 7r{z) and z{O) = y, then [-v, v] ~ 7r{z) and z{[-v, v]) ~ OJl.Y' It is suitable to pass to a new rectangular system of coordinates and to put its initial point at y, to direct the x-axis perpendicularly to the line 1, and to take as positive the direction from the point y to the line 1. Let a function Z E D{F,OTJY x ~) is represented in this system of coordinates by its coordinate functions z{t) = (Xl{t),X2{t»). Since x~(t) ~ p* here for all t E 7r(z), the function Xl{t) monotonically increases and is absolutely continuous (see Theorem 4.9.1). The same inequality x~ (t) ~ p* implies that its inverse function t = a(xl) is absolutely continuous. Define the function cpz : Xl(7r{Z) -- ~ by the formula cpAxd = x2{a{xl»)' By Remark 4.11.1 and Lemma 4.9.8 the function cpz is generalized absolutely continuous. Assertion 4.11.1 implies that (3.3)
, () cp s z
= x~{a{s» x~ (a{s»
for almost all s E 7r{CPz» = Xl{7r(Z». Notice that diamz O(y, 'fJ) < f!:; < 00. The condition D(F) E Ace(V) implies easily that H(D(F,OTJY x ~» E Rce(OTJY)' where the mapping H : D(F,OTJY x ~) -- Cs(OTJY) is defined by the formula H(z) = cpz. By Theorem 8.1.2 H(D(F,OTJY x~» E Rcek(OTJY)' By Lemma 8.1.2 the set Wl = {z: Z E D(F,OTJY x ~), 7r(z) = [0, vj, z(O) = y} is compact. The projection Q of the compactum {z(v): Z E 'lid in
488
CHAPTER 16
the x-axis does not contain the origin of coordinates. Therefore the number b = inf Q < TJ is positive. By Lemma 8.1.2 and Theorem 8.1.4 the set w2 = {ip: ip E H(D(F, Of/Y x JR)), 7l'(ip) = [0, b], ip(O) = O} is compact and connected. By Theorem 3.6.1 the set WI is equicontinuous. Therefore there exists a number v* E (0, v) such that if z E WI and s E [0, v*], then iiz(s) - yii < b. Let s E (0, v*), z E D(F), 7l'(z) 3 0, sand z(O) = y. Let z* E (D(F))-+ be an extension of the function z (existing by Lemma 6.6.2). By the choice of v and J1. we have [0, v] ~ 7l'(z*), z*([O, v]) ~ O/-iY' Therefore z*i[o,lI] E wt and the number b is not greater than the x-coordinate of the point z*(v). For some s* E [v*, v] the x-coordinate of the point z* (s*) is equal to b. Let ip* = H (z*i[o,s*])' Evidently z(s) E Gr(ip*) and ip* E w2. Thus
As = {z(s): z E wd = u{{z(s): H(z) = ip}: ip E wd = u{ {z(s): z E H1I(ip)}: ip E wd,
where HI = HiD(F,O'(Y,/-I)x[O,IIj)' The set D(F, O,(y, J1.) x [0, v]) is compact. The mapping H is continuous (see Theorem 3.3.1). Therefore the set W3 = Hll(w2) is compact and the mapping h : W2 --t As, h(ip) = {z(s): z E H1I(ip)}, is upper semicontinuous. By Theorem 2.6.4 our aim will be achieved when we show that for every ip E W2 the set h( ip) is connected. The measurability of the function ip', the Luzin's theorem 4.13.7, and the membership F* (JR x V) imply the measurability of the mapping G: 7l'(ip) --t JR2, G(s) = F(s,ip(s)) n {).ip'(S): A ~ O}. This easily implies the measurability of the function g(s) = P,iO,G(s)). By Lemmas 4.12.3 and 4.12.4 the mapping (3(s) = G(s) n O,(O,g(s)) is measurable. Notice that (3(s) is a point of G(s) nearest to 0. Take an arbitrary function z = (Xl' X2) E H1I(ip). By (3.1) and by the membership z E D(F) we have Zl(t) E G(XI(t)) for almost all t E [0, d] = 7l'(z). Therefore there exists a number p(t) E (0,1] such that (3(XI(t)) = p(t)Z'(t). Let for r E [0,1] Pr(t) = max{p(t),r}. Let (3(s) = ((3I(S),{32(S)). Then (31(XI(t)) = p(t)x~(t). Since
the equality p(t) = /31~~(g)) implies the measurability of the functions
p(t) = Po(t), Pr(t),
°< r
~ 1,
and the estimate Pr(t) ~ p* /x~ (t). This implies the Lebesgue integrability of the function l/Pr(t) (~ (p*)-IX~(t), see Theorem 4.9.1 and Lemma 4.7.6).
489
Two-dimensional systems. Let
t
Br(t) =
J
du
Pr(U) ,
o
Let the function C r : [0, dr ] --7 [0, d] be inverse to the function Br and zr(t) = z(Cr(t». By Lemmas 4.9.7 and 4.9.8 the function Zr is absolutely continuous (see also Theorem 4.9.1) and z~(t)
= ZI(Cr(t»C;(t) = zl(Cr(t))Pr(Cr(t))
E G(x~Cr(t»)
for almost all t E [O,d r] (see §4.11). Therefore Zr E D(F), Zr E Hll(cp). For 0 ~ rl ~ r2 ~ 1 we have
IIBr2 - Brl " ~
Jd(1 - () Prl t o
1)
- ( ) dt ~ Pr2 t
Jd x' (t) 0
- * dt P
< O.
By Lebesgue's theorem 4.7.4, the Corollary of Theorem 3.5.4, and Theorem 2.1.1 the mapping B*(r, t) = (r, Br(t», 0 ~ r ~ 1, 0 ~ t ~ d, is continuous and its inverse mapping
(B*)-l: {(r,t): 0 ~ r ~ 1, 0 ~ t ~ dr }
--7
[0,1] x [O,d]
is upper semicontinuous (see Corollary of Theorem 1.7.8). By virtue of the monotonicity of the functions Br the mapping (B*)-l is single valued. Therefore it is continuous in the usual sense. The formula qz(r) = zr(s) (= z(Cr(s» = z((B*)-l(r, s»)) defines a continuous mapping qz : [0,1] --7 Gr(cp). Evidently qz([O, 1]) ~ As. By Theorem 2.6.1 we will prove the connectedness of the set As = U{qz([O, 1]): z E Hll(cp)} when we establish the nonemptiness of the intersection n{qz([O, 1]): z E Hll(cp)}. Take arbitrary z, z* E Hll(cp) and show that qz(O) = qz. (0). Keep for z the notation of the previous reasoning. In the case of z* we will add to this notation the upper index *. We have z~(t) =
Z'(CO(t))p(Co(t)) = ,8(Xl(CO(t») = ,8(XOl(t»
for almost all t E [0, dol (= 7r(zo» and (z~)'(t) = ,8(X~l (t»
490
CHAPTER 16
for almost all t E
[O,d~]
(=
7r(z~)),
where zo(t) = (XOl(t),X02(t)) and
z~(t) = (X~1(t),X~2(t)).
Let for definiteness XOI (do) ~ X~l (d~). Then we can consider the absolutely continuous function (see Lemma 4.9.8) XO/X~l : [0, d~] ---+ [0, do]. We have (X~lnt) ( -1 * )/(t) X01X 01 = X01 I ( -l( X01 * (t)))' X01 By the above remarks we have in the numerator the first coordinate of the vector ,6(X~l (t)). We have in the denominator the first coordinate of the vector ,6(XOl(XO/(X~l(t)))) = ,6(XOl(t)). Therefore (XO/X~ly(t) = 1, z~(t)
=
(X~l(t),CP(X~l(t)))
XOIIX~l (t)
=
== t,
X~l (t)
(X~l(t),CP(XOl(t)))
== XOI (t),
= zo(t), z*(s) = zo(s).
The theorem is proved. • Lemma 3.1 (Lebesgue). Let"( be a cover of a compact subset K of a metric space (X, p) by sets being open in X. Then there exists a number ~ > 0 such that the ~-neighborhood of an arbitrary point of the set K lies in an element of the cover "(. Proof. By virtue of the compactness of the set K it is sufficient to consider only the case of a finite cover T "( = {G 1 , ••. , G j}. Assume that none of the sets G 1 , ... , G j coincides with X (in the opposite case the assertion is obvious). For x E X and i = 1, ... ,j put CPi(X) = p(x, X \ Gi ). The continuity of the functions CPl, ... ,CPj is established in Corollary of Lemma 2.3.1. This implies of the continuity of the function cp(x) = max{ CPl (x), .. . ,CPj(x)}. Since the family,,( covers the set K, cp(x) > 0 for every point x of this set. By Remark 2.3.5 inf cp(K) > O. If now we take an arbitrary number ~ E (0, inf cp(K)) then all imposed conditions become fulfilled. The lemma is proved. • Now we have the possibility of discussing the main topic of this section. By Theorem 3.1 under our restrictions on F we can apply all reasonings of the previous section to the space Z = D(F). Let f : 8 ---+ V be an arbitrary closed path free of stationary points of the space Z. For every point x E f(8) find a number 1](x) > 0 satisfying (3.2). For the cover {O(x,1](x)) : x E f(8)} of the compactum f(8) find ~ according to Lemma 3.1. Take an arbitrary a E (O,~). For s E 8 put ha(s) =
{II:II :
Ha(s)
U
E cc(F(O(f(s), a)))} ,
= n{[ha(O(s, e))]: e > O}}.
491
Two-dimensional systems.
The connectedness of the set cc( F( O(J (s), a))) and Corollary 1 of Theorem 2.6.4 imply the connectedness of the set ha(s). Find co > 0 such that 1(0(s, co)) ~ O(J(s), min{~, ~ -a}). If 0 < c < co and s* E O(s,c), then O(J(s*),a) ;2 O(J(s),~) and
ha(s*) ;2 F
(0 (J(s), ~)) =I
0.
By Theorem 2.6.1 the set ha(O(s, c)) = U{ha(s*) : s* E O(s, c)} is connected. Theorem 2.6.2 and Corollary of Theorem 3.3.3 imply the connectedness of the set Ha(s). On the the other hand, under the same choice of c and s· we have O(J(s·),a) ~ O(J(s),~). Therefore the set ha(O(s,c)) lies in the circle S in a segment of the length < 7r. Hence we can say it about Ha(s) too. The nonemptiness of the set Ha(s) is obvious. Thus Ha(s) is a proper connected compact subset of the circle S. Let a sequence {Sk: k = 1,2, ... } ~ S converge to a point s. By remarks of Example 1.5.3 lim top[h( O(s, 2p(Sk, s)))] = Ha(s). k-+oo
By Theorem 1.7.2 the sequence {[h(O(s, 2P(Sk' s)))] : k = 1,2, ... } converges to the set Ha(s) in the sense of condition (1.7.1). Since Ha(Sk) ~ [h(O(s, 2P(Sk' s)))] we have the convergence in the same sense of the sequence {Ha(Sk) : k = 1,2, ... } to the set Ha(s). By Corollary of Lemma 1.7.1 this implies the upper semicontinuity of the mapping Ha. Thus Ha E e(S, S) and we can apply the theory of §2.8 to the mapping
Ha· Let
z
E
D(F),
Im(z)
al < 0 < bl ,
7r(z) = [al' bd,
~ 0, (J(s),~),
z(al),z(bd
E
S
z(O) = 1(s),
(1(S),~).
By Theorem 4.10.1
z{ad - z(O) z(bd - z(O) , b al 1
E cc
('([ b I)) z al, 1
(F (0, (J( s), ~) ) )
~
CC
~
cc(F(O(J(s), a))),
492
CHAPTER 16
With the notation of §2 {
u -
I(s)
}
Ilu - l(s)11 : u E PI ~
u - I(s) { Ilu _ l(s)11
-He.(s),
:u
E P2
}
~ He.(s).
Therefore P(f(S),~) ~ He.(s). By Lemma 2.5.1 and Theorem 2.8.1 the degree of the mapping He. is equal to j(Z, I). For s E S put Ho(s) = n{He.(s) : < a < O. The Corollary of Theorem 3.3.3, the compactness of the circle S (which implies the fulfilment of condition (1.6.3)) and Theorem 2.5.1 imply the membership Ho E e(S, S). Thus we can apply the theory of §2.8 to the mapping Ho. By Lemma 2.8.2 and Theorem 2.8.1 the degree of the mapping Ho is equal to j(Z, I). We did not define the mapping Ho directly by the right hand side of our inclusion and we have used auxiliary constructions. Let us simplify the description of the set Ho(s) under the additional assumption of the nonemptiness and of the compactness of the set F(f(s)). Take an arbitrary 8 E (0,00). By virtue of the upper semicontinuity of the mapping F at the point I(s) (see Theorem 2.5.2) there exists a number a* E (0,0 such that F(O(f(s), 2a*)) ~ 0(F(f(s)),8). By virtue of the continuity of 1 for some c* > we have I(O(s,c*)) ~ O(f(s),a*). For s* E O(s, c*) we have
°
°
O(f(s*), a*)
~
O(f(s), 2a*),
F(O(f(s*), a*))
~
O(F(f(s)), 8).
Since the last set is convex (see Lemma 2.4.3),
he.< (s*) =
~
{II:II :
u E cc(F(O(f(s*), a*))) }
{II:II : u
E
[O(F(f(s)), 8)]}
by Lemma 2.4.5. In view of the arbitrariness of s* E O(s, c*)
he.< (O(s, c*)) Ho(s)
~
{II:II : u E [O(F(f(s)), 8)]} ,
~ He.< (s) ~ {II:II :
u E
[O(F(f(s)), 8)]} .
Since the number 6 E (0,00) is taken arbitrarily, by virtue of the compactness of the set [0(F(f(s)),6)] (see §2.4)
Ho(s)
~ n { {II:II =
{II:II :
:
u E
u E
[O(F(f(s)), 6)]} : 6 E
F(f(s»} .
(o,oo)}
493
Two-dimensional systems.
The last set lies in H 0 (s ). Therefore
Ho(s)
=
{II:II :u
E
F(f(s))}.
If the mapping F is single valued (Le., if we deal with a differential equation) then the last equality goes into the equality
F(f(s)) Ho(s) = IIF(f(s))II. This equality is usually used in the definition of the index for equations with continuous right hand sides (see [HaJ). Notice a situation when arguments of the last two sections allow us to avoid a complication of the calculation in the comparison with classical cases and to expand the index theory to equations with singularities. Let Z E Acek(V) (methods of the verification of this condition are developed) and f : 8 - t V be a closed path. Assume that in a neighborhood W of the set f(8) the space Z coincides with the solution space of the equation y' = F(y): ZRxW = D(F, ~ x W). We may calculate the value j (Z, f) with the use of F as this is made in the framework of the classical theory and then we may apply the result of the calculation to estimate the existence of stationary points of the space Z, without caring about a description of singularities.
4. Poincare-Bendixson theorem Let V be an open subset of the plane, (Z, Zoo) E B*(V) and Zoo E Ak(V). Lemma 4.1. Let K ~ V be a closed disc. Let x be its center and diamz"" K < 00. Then there are disjoint intervals SI and S2 of the circle 8 bounding the disc K, a number T E ~, and a neighborhood Ox of the point x such that if Z E Z, Im(z) ~ K, Im(z) n Ox # 0, 7r(z) = [c, d], c ~ T and z(c), z(d) E 8, then z(c) E SI and z(d) E S2. Proof. Construct segments PI and P2 according to the reasoning of the beginning part of §2 for the space Zoo. Take as SI and S2 arbitrary disjoint intervals of the circle 8 such that the first interval contains the segment PI and the second interval contains the segment P2. Show that under this choice the imposed conditions hold. Assume the opposite. Then for every k = 1,2, ... there exists a function Zk E Z such that 7r(zd = [Ck' dkj, Ck > dk , Im(zk) ~ K, Im(z)nO(x, # 0, Zk(Ck), zk(dk ) E 8 and either the point Zk(Ck) does not belong to the interval SI, or the point zk(dd does not belong to the interval S2. Lemma 15.2.3 and the finiteness of the diameter of the set K with respect to the space Zoo imply the boundedness in. the totality of the number
i)
494
CHAPTER 16
dk - Ck, k = 1,2, .... The same lemma implies now the existence of a function z E Zoo such that Im(z) ~ K, x E Im(z) and either z(inf 7r(z)) FI. 81, or x(SUP7r(Z)) FI. 82. This contradicts the choice of the intervals 81 2 PI and 82 2 P2· The lemma is proved. • Lemma 4.2. Let z E Z+ and 7r(z) = [a, 00). Let the curve z(t) have no self-intersections. Let x be an arbitrary point of the set 0+ (z) n V. Let the point x not be a stationary point of the space Zoo. Then there are a neighborhood Ox ~ V of the point x and a function Zo E Zoo such that Ox n O+(z) = Ox n Im(zo). Proof. By remarks of §15.1 for some c > 0 the diameter of the set K = Of (x, c) ~ V with respect to the space Zoo is finite. Find Ox, T > a, 81 and 82 according to Lemma 4.1. A passage from a given neighborhood Ox to a smaller one cannot lead to the non-fulfilment of the condition imposed. Therefore we can require in addition, that Ox = O(x, 8) for some
8E(O,c). The set Z-I(O(X, c)) is open in the half interval 7r(z). By the Corollary of Theorem 2.6.8 it falls into the union of the countable family 'Yo of its connected components. Let 'Yl be the subfamily of 'Yo consisting of all elements of 'Yo, which lie on the right of T and which do not meet the set Z-I(OX). For each element G of the family 'Yl select a point t(G) such that z(t(G)) E Ox. The set {t(G): G E 'Yd has no limit points. (Between every two points of this set there is a point in which the function z takes value in the set V \ O(x, c). Therefore if a limit point exists the value of the function z at it must belong both to the set V \ O(x, c) and to the set Of (x, 8). This is impossible because these sets are disjoint.) This implies that the points t(G), G E 'Yl, may be enumerated in the order of their increasing: tl < t2 < t3 < .... Enumerate in the corresponding way elements of the family 'Yl: 'Yl = {(ak, bk ) : k = 1,2, ... }, where tk = t((ak, bk )). We have z(ak) E 81 and z(b k ) E 82. Take an arbitrary number k = 1,2, .... The points z(ad, z(ak+1) are different and they define an order in the interval 81. It is convenient to use the usual terminology to describe this order. Assume for definiteness that the point z(ak+1) lies on the left of the point z(ak). Show that the point z(ak+2) lies in 81 on the left of the point z(ak+1). In order to do this joint the curves z(t), ak < t < bk, and z(t), ak+1 < t < bk+1, by a segment I lying entirely in the set Ox (let z(t*), t* E (ak, bk ), and z(t**), t** E (ak+1, bk+1), be the endpoints of this segment), such that the closed path>' composed by the segment I and the curve z(t), t* < t < t**, has no self-intersections (Figure 16.8). Let the region Vi be bounded by the path composed by the segment I, by the curve z(t), ak+1 ~ t ~ t**, by the interval with the endpoints z(ak) and z(ak+1)' which lies in the interval 81, and by the curve z(t), ak ~ t ~ t*. Let the region V2 be bounded by
Two-dimensional systems.
495
the path composed by the segment l, by the curve z(t), t* ~ t ~ bk , by the interval with the endpoints z(b k ), z(b k +1) lying in the interval 82, and by the curve z(t), t** ~ t ~ bk+1. The regions cannot lie one in another, because the boundary of every of them contains elements which do not lie in another region, namely the corresponding part of the arcs 81 and 82. Therefore the regions V1 and V2 adjoint from different side the segment l. Hence they lie in different components of the complement to the path A. The boundary of Vi contains the point z(ak). The boundary of V2 contains the point z(bk+d. Thus the points z(ak) and z(bk+d lie in different components of the complement to the path A.
. ---- ~ z(ak+-rl
.-
81
,/"
,_ ...-_._--------
v:
1/ .~
._-------
z(ak)
.....
..
...
Figure 16.8
Assume now that the point z(ak+2) lies in 81 to the right of the point z(ak+d· By the above remarks the curve z(t), bk+1 ~ t ~ ak+2, meets the path A. Since the curve z(t) has no self-intersections the curve z(t), bk+1 ~ t ~ ak+2, meets the segment l. Hence it meet the set Ox. This contradicts the choice of the point ak+2. We have obtained a contradiction, which gives what was required. Thus the points z(ak), k = 1,2, ... , constitute a monotone sequence in the interval 81. Since the curve z(t) has no self-intersections the points z(b k ), k = 1,2, ... , are ordered in the interval 82 in a corresponding way.
496
CHAPTER 16
For k = 1,2, ... and t E [0, bk - ak] put z;; (t) = z(t + ad. Lemma 15.2.3 and the finiteness of the diameter of the set K with respect to the space Zoo imply that the sequence {z;; : k = 1, 2, ... } has a subsequence {z;; : k E A} converging to a function Zo E ZOO' For k = 1,2, ... let Fk denote the closure of the region whose boundary is a path composed by the curve z;; (t), 0 ~ t ~ bk - ak, and by the arc of the circle S with the endpoints z;;(O) and z;;(b k - ad, which does not contain the points Zk+l(O) and zk+l(bk+l - ak+l)' The curve zk(t), 0 < t < bk - ak, goes in the interior of the compactum Fk+l and
Therefore
limtopsupz([ak,bd) = lim top sup Im(zZ) k~oo
k~oo
= [U{Fk : k = 1,2, ... }] \ (U{Fk : k = 1,2, ... })
[U{Fk : k E A}] \ (U{Fk : k E A}) = lim top sup Im(zZ) = Im(zo). X
k~oo
kEA
By the choice of the points ak and bk , k = 1,2, ... ,
O+(z) nox = limtopinfz([ak,bd) nox. k~oo
Therefore O+(z) n Ox = Im(zo) n Ox. The lemma is proved. • Theorem 4.1. Under the hypotheses of Lemma 4.2 let the set O+(z) be nonempty and compact, lie in V, and have no stationary points of the space Zoo. Then there exists a function Zl E Zoo such that O+(z) = Im(zl) and Zl (t) is a simple loop. Proof. Denote by M the set of all functions z* E Zoo such that Im(z*) ~ O+(z) and the curve z*(t), t E (7r(z*)), has no self-intersections. Let 0: = sup{I7l"(z): z EM}. The passage from the neighborhood pointed in Lemma 4.2 to a smaller one cannot lead to the non-fulfilment of the stated conditions. So we can require in addition that diamzoo [Ox] < 00 (however, it is already guaranteed by the constructions in the proof of Lemma 4.2). In this case by Lemma 15.1.3 when we shorten, if necessary, the domain of the function Zo of Lemma 4.2, we obtain the fulfilment of the condition Zo E M, which implies the estimate 0: > O. In addition, the mapping
Two-dimensional systems.
497
turns out to be an homeomorphism (see Theorem 2.2.1). Hence every point E O+(z) possesses a neighborhood Gx in O+(z) homeomorphic to an interval of the real line and diamzoo Gx < 00. By virtue of the assumed compactness of the set O+(z) its open cover {Gx: x E O+(z)} has a finite sub cover {GXl,'" ,Gxp}. Evidently for every function z* E M and for every set Gx, x E O+(z):
x
either Im(z*) or Gx
~
~
GXj
Im(z*) and the set (Z*)-l(GX) is an interval lying in 7r(z*)j
or the set (z*) -1 ( G x) is a half interval or falls into the union of two disjoint half-intervals being open in the segment 7r(z*). Therefore a ~ 2 E{diamzoo GXi: i = 1, ... ,p} < 00. The definition of a and the autonomy of the space Zoo imply the existence of a sequence {zZ: k = 1,2, ... } ~ M, 7r(zZ) = [0, ak], ak - t a. By virtue of the condition Zoo E Ac(V), of the compactness of the set O+(z) and of the estimate 17r(zZ) ~ a we can assume in addition that the sequence {Zk: k = 1,2, ... } converges to a function Zl E Zoo' By virtue of the continuity of the mapping 7r we obtain 7r(zd = [0, a]. In addition, Im(zl) ~ O+(z). Show that the mapping zll(7r(zt)) is injective. Assume the opposite, < tl < t2 < a. Then we can select i.e., that zl(td = Zl(t 2), where points al E (O,td and b2 E (t2,a) such that zl([al,td) ~ Gzl(td and
°
Zl([t 2,b2])
~
GZl(t l ). The set G Zl (td is homeomorphic to an interval. Use this fact to describe geometric particularities of situations arising. The points Zl (al) and Zl (b 2 ) lie in the interval G Zl (t l ) on different sides of the point Zl (t l ), because in the opposite case Lemma 15.1.3 easily implies the estimate diamzoo Gzl(td = 00. Likewise there are points bl < a2 of the segment [tl' t2] such that zd[t l , bd) = Zl([t 2, b2]) and zl([al, td) = zl([a2, t2])' By virtue of the convergence zk - t Zl for a large k the value Zl (t l ) is taken by the function zZ at least twice (on the segments [aI' bl ] and [a2' b2]). This contradicts the condition zZ E M. So Zl E M. The sets Gx, x E O+(z), cover in their totality the compactum O+(z). They are homeomorphic to intervals of the real line. Thus functions from M are either simple paths or simple loops. Of course, this is true for the function Zl too. For Zl the first case is not possible, because here we have the possibility of continuing the function Zl over 0 or over a with the keeping of the belonging to M (see Lemma 4.2). Thus Zl is a simple loop. Likewise the set Im(zd is open in O+(z). Its closedness in
498
CHAPTER 16
n+(Z) is obvious (see Theorem 1.7.8). Thus the set Im(zl) is a connected component of the set n+(z). By Lemma 15.2.2 the set n+(z) is connected. Thus n+(z) = Im(zl)' The theorem is proved. • Corollary. Let Zl E Acek(V), z E 7l'(z) = [a,oo). Let the curve z(t) have no self-intersections. Let the set n+(z) be compact, lie in V, and have no stationary points of the space Zl' Then there exists a function Zl E Zl such that n+(z) = Im(zl) and Zl (t) is a simple loop. To obtain the corollary it is sufficient to refer to remarks of Example 15.1.2. • When we take as the space Zl the solution space of an ordinary differential equation with a continuous right hand side (often under the additional assumption of the fulfilment of condition (u)) the Corollary of Theorem 4.1 is just the Poincare-Bendixson theorem. Our Theorem 4.1 is its generalized version. It may be applied in an essentially broader case even if we restrict the consideration to equations with continuous right hand sides. The remark of §15.1 helps the reader to give corresponding examples without essential efforts. Of course, for every such an example we can give a list of restrictions on right hand side of the equation, which allows our constructions (or constructions of other proofs of Poincare-Bendixson theorem), and we may obtain a theorem in a spirit of the classical theory. But in this way we shorten immediately the domain of the applications of the result. Our methods allow us to apply Theorem 4.1 and its Corollary (as other our results) without essential additional efforts in many situations far from those of the classical theory. The function Zl occurring in Theorem 4.1 may be extended over the entire real line by the periodicity. Therefore we obtain: Theorem 4.1 *. Under hypotheses of Theorem 4.1 there exists a periodic function Zl E Z;:;;/ such that n+(z) = Im(zd and if p is a (smallest) period of the function Zl, then Zl (td 1= Zl (t 2) for 0 :::;; h < t2 < p. • We may reword Corollary of Theorem 4.1 in an analogous way too. Besides the function Zl from Theorem 4.1 *, every function that is a translation of Zl in t, i.e., is defined by the formula z* (t) = Zl (t - to) (for some to E IR), also satisfies the stated conditions. Theorem 4.2. Under the hypotheses of Theorem 4.1* let the function Zl be defined uniquely up to translations in t, {ak: k = 1,2, ... } ~ [a, 00), ak - t 00. For k = 1,2, ... and t E [0,00) let zZ(t) = z(t + ad. Let z(ak) = zZ(O) - t Zl(O). Then for every segment I ~ [0,00) the sequence {ZZ II: k = 1,2, ... } converges uniformly to the function zllI' Proof. I. It is sufficient to prove the. assertion under the additional assumption inf I = O. Do it.
zt,
Two-dimensional systems.
499
II. Denote by b the upper bound of right endpoints of segments, for which the mentioned condition holds. When b = 00 we have what was required. III. Let b < 00. Lemma 7.2.1 implies that the segment [0, b] satisfies also our condition. The definition of b implies that no larger segment satisfies this condition. This means that if we change the sequence {ak : k = 1,2, ... } to the sequence {ak + b: k = 1,2, ... } then there is no segment I of nonzero length satisfying the stated condition. Thus the problem reduces to the proof of the existence of a segment I of nonzero length with the initial point at 0 such that the sequence {Zk II : k = 1,2, ... } converges uniformly to the function ziII. For x = Zl (0) find c > 0 such that the set O(x, 2c) lies in the set Ox of Lemma 4.2 and diamzoo O(x, 2c) < 00. For Zoo, to = 0, Yo = Zl (0) = x and U = ~ x O(x, c) find JL > v > 0 according to Theorem 6.7.3. By the assumptions made and Lemma 7.2.1 the sequence a = {zkl[o,lI] : k = 1,2, ... } contains a subsequence converging to a function Z2 E Zoo. Assume that it differs from zll [0,11]. By the choice of v we have Im( Z2) ~ O( x, c). Since diamzoo O(x, c) < 00, Z2(t) is a simple path (see Lemma 15.1.3). Take now arbitrary tl E ~ such that Zl(t l ) = ZI(V). Put t2 = inf{t: t > t l , Zl (t) = x}. The function for t E [0, v], for t E [v, v + t2 - td is defined on the segment J = [0, v + t2 - td. It belongs to the space Zoo and Z3 =I zlIJ· Evidently it satisfies all conditions imposed in Theorem 4.1 on the function Zl (see the proof of Theorem 4.1). Its extension by periodicity to the entire real line cannot be obtained from the function Zl by a translation in t. The contradiction obtained implies the convergence of the sequence a to zll [0,11]. The theorem is proved. • Now make an important addition to Theorem 4.2 in relation to the condition of the uniqueness of the function Zl which occurs there. Show that if Zoo is the solution space of an equation y' = f(y) with the continuous (it is sufficient, on the set O+(z)) function f then the condition of uniqueness holds. Notice that the absence of stationary points of the space Zoo in the set O+(z) is equipotent in this case to the situation when the function f does not vanish on O+(z). Let ZI,Z2 E Zoo, 7r(zd = 7r(Z2) = [0,00), Zl(O) = Z2(0) and Im(zd = Im(z2) = O+(z). Let M denote the set of all numbers t E [0,00), for which we can define a continuous increasing function gt : [0, t] -+ [0, 00) such that g(O) = and Zl(S) = Z2(gt(S)) for every value of the argument S E [0, t]. The set M nonempty: it contains, for instance, the point 0. The
°
500
CHAPTER 16
above description of the set O+(z) implies that the set M is open. The continuity of the functions Zl and Z2 implies the closedness of the set M and the coincidence of the functions gt, and gt2 on the intersection of their domains. The connectedness of the half interval [0,00) now implies the equality M = [0,00). The previous remark implies that the formula gl[o,tj = gt defines uniquely a continuous non-increasing function 9 : [0, 00) ~ [0,00). By remarks of §8.5 the functions Zl and Z2 are continuously differentiable and satisfy the equation for every value of the argument. When we pass to the limit as 8 ~ 00 in the equality
Z2(g(8 + 8)) - Z2(g«8)) g(8 g(8+8)-g(8) .
+ 8) -
g(8)
8
by virtue of the non-vanishing of the derivatives of the functions Zl and Z2, we obtain the existence of the derivative of the function 9 and the equality J(Zl(8)) = J(Z2(g(8)))g'(z). This implies the equality g'(8) == 1. Here g(O) = 0 and therefore g(8) == 8, Zl = Z2' Thus the needed condition holds.
5. Neighborhoods of stationary points in the plane In this section we generalize the main theorems of Chapter VIII of [HaJ. Let auto-homeomorphisms 6, of the real line be associated with positive A and (5.1)
6,(0) = 0,
(5.2)
6,(t l ) < 6,(t 2) for every tl < t2 and A > O.
For t E IR and y E IRn put 0), U = IR x V, Vo = V \ {O} and Uo = IR x Vo. For T E IR and (t, y) E U put 'liAt, y) = (t - T, y). Consider the following conditions: (5.3) ( is a compact subset of the space Rce(U), the function to every space Z-+, where Z E (j (5.4)
~T(Z) E ( for every Z E ( and
T
Zo
== 0 belongs
~ OJ
(5.5) {6. : A > O} is a family of auto-homeomorphisms of the real line satisfying conditions (5.1), (5.2) and O.
Two-dimensional systems.
501
For an arbitrary 0 E IR denote by l(O) the ray {>.(cosO,sinO): >. > O}. For 01 < O2, O2 - 01 < 27r denote by S(Ol,02) the set U{l(O): 01 < 0 < 02}. A vector u( cp) = (cos cP, sin cp) is called non-characteristic ifthere exists a number 8 E (0, 7r) such that if CPl, CP2 E IR and cp-8 < CPl < cP < CP2 < cp+8, then either:
zt
(5.6) the set {z: z E ZUo U o' Im(z) ~ S(CP1,CP2), Z E 0 is empty and if z E U{Z-+: Z E 0, c E 7r(z), >. > 0, z(c) = >.u(cp) then for some a, bE 7r(z) we have c E [a, b], z([a, b]) ~ [S(CP1' CP2)j, z(a) E l(cpd and z (b) E l (CP2);
or:
zt
(5.7) the set {z: z E ZUo U o' Im(z) ~ S(CP1, CP2), Z E 0 is empty and if z E U{Z-+: Z E 0, c E 7r(z), >. > 0, z(c) = >.u(cp) then for some a,b E 7r(z) we have c E [a,bj, z([a,b]) ~ [S(CP1,CP2)]' z(a) E l(CP2) and z(b) E l(cpd. A vector u( cp) = (cos cP, sin cp) is called characteristic if it is not noncharacteristic. By (5.5) for a non-characteristic vector u( cp) either for all points of the ray l (cp) (being rather close to 0) condition (5.6) holds, or for all such points condition (5.7) holds. This implies, in particular, that all such points are points of strong entry (see [HaJ) for S (cp - 8, cp) and are points of strong exit for S( cP, cP + 8), or vice versa. Denote by ~ the set of all cP E IR such that the vector u( cp) is noncharacteristic. Consider the following situation: (5.8)
X E R(U), x E X+, 7r(x) = [0,00);
(5.9) a function", : [0,00) -+ (0,00) is continuous, the set P ~ (0,1/) oflimit points of the expression ",(t)lIx(t)11 as t -+ 00 is nonempty and compact, the generalized sequence of the spaces {-i>e'1(~) ~AX): T > O} converges to the set ( as T -+ 00; (5.10) 0: [0,00)
-+
IR is a continuous function and
x(t) . IIx(t)1I = (cos O(t), smO(t)),
0= liminfO(t), D = limsupO(t) t-+oo
t-+oo
Theorem 5.1. Let conditions (5.3 - 5), (5.8 - 10) hold. Then (0, D) nE = 0.
502
CHAPTER 16
Proof. I. Take an arbitrary Jl. E (0,2- 1 inf P). By our assumptions there exists to > 0 such that TJ(t)lIx(t)11 E 0ttP for all t ~ to. II. Assume that our assertion is false. Then there exists
C <
X«Sk' t k)) ~ S(
X«S2m+l,P2m+l)) ~ S(
and The convergence in (5.9), the first parts of conditions (5.6-7) and Theorem 7.2.2 imply the boundedness in the totality of lll"(zd, k = 1,2, .... By (5.9) the sequence {Z2m+l: m = 1,2, ... } contains a subsequence converging to a function Zl E U(, 7r(Zl) = [Cl,O], and the sequence {Z2m: m = 1,2, ... } contains a subsequence converging to a function Z2 E U(, 7r(Z2) = [0, C2J. Here (for Zl) either IIZl(cdll E 8(02ttP), or zl(cd E l(
503
Two-dimensional systems.
a) limt-+oo O(t) = 00; or
b) limt-+oo O(t) = -00; or c) -00
~
D
•
< 00.
We have also: Theorem 5.3. Let conditions (5.3 - 5), (5.8 - 10) hold and [E] = JR. Then: either a) limt-+oo O(t) = 00; or
b) limt-+oo O(t) = -00; or
c) the limit limt-+oo O(t) E JR \ E exists. Proof. It remains only to analyze the case c) of Theorem 5.2. The condition [E] = JR and Theorem 5.1 imply the existence of the finite limit limt-+oo O(t) = 00 , It remains to show that 00 E JR \ E. This follows immedi• ately from the definition of E. The theorem is proved. Consider a particular case. Let Z E Ace(U) and {6: ). > O} be a family {6: ). > O} of auto-homeomorphisms of the real line satisfying (5.1) and (5.2). Let ~e~ (Z) ~ Z for every). > O. In this case the family ( = {Z} satisfies (5.3-5) and Theorem 5.3 implies: Corollary. Let X E Ace(U), x E X+, 7l'(x) = [0,00), limt-+oo Ilx(t)1I = 0, ~eJX) --t Z as ). --t 00. Let 0 : [0,00) --t JR be a continuous function and x(t)lIx(t)II-1 = (cosO(t),sinO(t». Then there exists the finite or infinite limit limt-+oo O(t). If the limit 00 = limt-+oo O(t) is finite, then 00 E 1R \ E .• Example 5.1. Consider the system {
X'
y'
+ p(x, y) = Q(x, y) + q(x, y), = P(x, y)
where
(5.11) P and Q are homogeneous polynomials of the degree m
~
1,
504
CHAPTER 16
(5.12) p,q are continuous andp(x,y),q(x,y) = o(rm), where r2 = x 2 +y2. Such a system is studied in §VIII.4 of [HaJ. In order to deduce Theorem VIII.4.1 of [HaJ from our assertion we may put ~A(t) = ).m-It. The change of variables
+ ).mp().-lx, ).-ly), y' = Q(x, y) + ).mq(). -lX, ). -ly).
X' {
= P(x, y)
Now (5.12) implies the convergence X()') denotes the solution space of the system {
X'
---7
Z as ).
---7
00, where Z
= P(x, y),
y' = Q(x,y).
This allows us to apply Corollary of Theorem 5.3, which gives what was required. Example 5.2. Keep the assumptions of the previous example in relation with P, Q and p. Let the number m is even and Q(x, 0) > 0 for X =I O. The system X' = P(x, y) + p(x, y!; { y' = Q(x,y) + c 1v1x may be studied analogously to the system of the previous example with a the small difference that we cannot use classical results on the continuity of the dependence of solutions on parameters to prove the convergence X()') ---7 Z. However, classical results are sufficient to prove this convergence outside the x-axis. Then we obtain the convergence in the entire plane with the help of arguments of Chapter 7. Example 5.3. Keep the assumptions of the previous example. The system X' = P(x, y) + p(x, y), { y' = Q(x, y) - axme-lvlx-2 (cos 2 tt+! may be studied analogously to the system of the previous example. The difference is related with the continuity of the right hand side. So the new system turns out to be closer to the domain of the classical theory; however, this case is not covered by classical assertions and it seems to be difficult to obtain corresponding results that are simpler, than those obtained with the use of the arguments mentioned. Example 5.4. Consider a two-dimensional equation y' = Ay + B(t)y, where A denotes a constant matrix and elements of the matrix B(t) are Lebesgue integrable on the interval (0,00).
Two-dimensional systems.
505
In this case take as ( the family consisting of one space Z of solutions of the equation y' = Ay and put ~(t) = t for every ,x, TJ(t) = Ilx(t)ll- l . The space {,,(~) WAX) coincides with the solution space of the equation y' = Ay + B(t + r)y and the convergence in (5.9) follows from remarks of Chapter 7. Now use Theorem 15.3.2 to obtain a generalized version of Theorem VIII.4.2 in [Ha]. Theorem 5.4. Let Z E Ace (V), a function Zo 1= 0 belong to Z, condition (5.5) hold for the family ( = {Z}, X E Acek(V),
(5.13) (~(X))uo
-t
Z as ,x
-t
00,
(5.14) every number 0 E [0 1 , Oo! satisfies (5.6) and every number 0 E [0 0 , O2] satisfies (5.7),
or (5.15) every number 0 E [0 1 ,00 ] satisfies (5.7) and every number 0 E [O O,02! satisfies (5.6).
Then for every /11 E [0 1 ,00 ] and /12 E [0 0 , O2 ] there exists a number P E (0,11) such that for every P E (0, Po) there exists a function x E X-UX+ such that 0 E 7r(x), Im(x) ~ 8(/11, /12)' Ilx(O)11 = p. Proof. I. The change of variables (t,y) - t (-t,y) transforms (5.14) to (5.15) and vice versa. Thus it is sufficient to prove the assertion in one of these two cases only. Assume for definiteness the fulfilment of (5.15). II. Fix arbitrary numbers /111 E (/11, ( 0 ) and /121 E (00 , /12)' Find for them positive 6 according to the conditions in the definition of a noncharacteristic vector. A passage to a smaller (positive) 6 does not break the fulfilment of these conditions. Therefore we can assume in addition that 26 < min{I/111 -/111,1/111 - 00 1, 1/121 -/121,1/121 - Ool}· III. Let
E
= 11/2. Show that there exists a number Eo E (0, E) such that
if i = 1,2, Z E Z-+, 0 E 7r(z), z(O) E 1(/1i) and IIz(O)11 ~ Eo, then there are numbers a E (-00,0] n 7r(z) and b E [0, (0) n 7r(z) such that z(a) E 1(/1i-(-1)i6), bE 1(/1i+(-1)i6), z([a,b]) ~ [8(/1i-6,/1i+6)]nO"o. If we assume the opposite then by the definition of a non-characteristic vector for every k = 1,2, ... there exists a function Zk E Z such that:
0, tk E 7r(zd, IIZk(O)1I
= E, IIZk(tk)1I
Im(zk) ~ [8(/1i - 6, J.Li
E (0,2- k ),
+ 6)] n OeD.
CHAPTER 16
506
Now Assertion 10.4.1 implies the existence of a function Z E ZUo U Z60 for some Z E ( such that E 7r(z), Im(z) ~ [S(J.Li - 8, J.Li + 8)] n [0,0'], Ilz(O) II = c. This contradicts the conditions in the definition of a noncharacteristic vector. IV. Now III and (5.13) imply the existence of a number AO ~ 1 such that for every A ~ Ao the space Xl = .x(X) satisfies the condition
°
if i = 1,2, x E Xl, c = SUp7r(z), z(c) E [(J.Li) and Ilz(O)11 ~ co/2, then there exists a number a E (-00,0] n7r(z) such that z(a) E [S(J.LI,J.L2)] and z([a, cD ~ S(J.LI - 8, J.L2 + 8) n 0,0'. So the space X and the number
CI
= c/ A satisfy the condition:
if i = 1,2, x E X-, c = SUp7r(z), z(c) E [(J.Li) and Ilz(O)11 ~ cd2 then there exists a number a E (-00,0] n 7r(z) such that z(a) E [S(J.LI, J.L2)] and z([a, cD ~ S(J.Li - 8, J.Li + 8) n 0,0'.
.
So we are about in the same geometric situation as in the proof of Theorem VIII.4.2 in [Hal. Referring to Theorem 15.3.2 completes the ~~
REFERENCES
[AI]
Aleksandrov P.S., Introduction to set theory and general topology, Nauka, Moscow, 1977 (Russian).
[Ar]
Artstein Z., The limiting equations of nonautonomuous differential equations, J. of Differential Equations, 25 (1977), 184-202.
[Bo]
Bokstein M.F., Theorems on the existence and uniqueness of solutions of systems of ordinary differential equations, Uchen. Zap. MGU, 15, N1 (1939), 3-72 (Russian).
[CL]
Coddington E., Levinson N., The theory of Ordinary Differential Equations, McGraw-Hill, New York, 1955.
[Da]
Davy J.L., Properties of solution set of a generalized differentional equation, Bull. Austral. Math. So., 6, N3 (1972), 379-398.
[Di]
Dieudonne J., Foundations of Modern Analysis, New York - London, Academic Press, 1960.
[Ed]
Edwards R.E., Functional analysis, Holt, Rinehart and Winston, New York, 1965.
[En]
Engelking R., General topology, PWN, Warszawa, 1977.
[ER]
Efimov N.V., Rosendorn E.R., Linear Algebra and Multidimensional Geometry, Nauka, Moscow, 1970 (Russian).
[Faf]
Filippov A.F., Differential equations with discontinuous right hand sides, Math. Appl. (Soviet Ser.) 18, Kluwer, Dordrecht, 1988.
[FF]
Fedorchuk V.V., Filippov V.V., General topology: Basic constructions, Moscow University Publishers, Moscow, 1988 (Russian).
[Fi]
Fichtengolz G. M., Course of differential and integral calculus, 3, Fizmatgiz, Moscow, 1963 (Russian).
[Fvv1)
Filippov V.V., Solution spaces of ordinary differential equations, Moscow University Publishers, Moscow, 1993 (Russian).
508
[Fvv2]
Filippov V.V., Ordinary differential equations: topological approach to the theory (to appear in Russian).
[Fvv3]
Filippov V.V., Topological structure of solution spaces of ordinary differential equations, Uspehi Mat. Nauk, 48, N1 (1993), 103-154 (Russian).
[Fvv4]
Filippov V.V., Basic topological structures of the theory of ordinary differential equations, Topology in nonlinear Analysis, Banach center publications, 35 (1996), 171-192.
[Fvv5]
Filippov V.V., On the acyclicity of solution set of ordinary differential equations, Doklady RAN, 352, N1 (1997) (Russian).
[Fvv6]
Filippov V.V., On Aronszajn theorem, Differenc. Umvnen., 33, N1 (1997), 11-15 (Russian).
[GOl]
Gossez J.-P., Omari. P., Nonresonance with respect to the FuCik spectrum for periodic solutions of second order ordinary differential equations, Nonlinear Anal. Th. Methods Appl., 14 (1990), 1079-1104.
[G02]
Gossez J.-P., Omari. P., Periodic solutions of a second order ordinary differential equation: a necessary and sufficient condition for nonresonance, Journal of Differential Equations, 94, N1 (1991), 67-82.
[Go]
G6rniewicz L., Topological approach to differential inclusions, Topological methods in differential equations and inclusions, 129-190, Kluwer, Dordrecht, 1994.
[Ha]
Hartman Ph., Ordinary differential equations, Wiley, New York, 1964.
[HW]
Hilton P.J., Wylie S. Homology theory, University Press, Cambridge, 1960.
[HZ1]
Habets. P., Zanolin F., Upper and lower solutions for a generalized Emden-Fowler equation, J. Math. Anal. Appl.,
[HZ2]
Habets. P., Zanolin F., Positive solutions for a class of singular boundary value problems, Preprint of Institut de Math. pure et appl., Universite Cath. de Louvain, Recherches de Math., 35, Oct.1993.
[KI]
Kluczny Cz., Sur certaines familles de courbes en relation aves la tMorie des equations differentielles ordinaires, Ann. univ. M.Curie-Sklodowska. Sec. A.Math., XV (1961), 13-40, XVI (1962), 5-8.
[Ku1]
Kuratowski K., Topology, 1, Academic Press, New York - London, PWN, Warszawa, 1966.
[Ku2]
Kuratowski K., Sur l'espace des fonctions partielles, Ann. di Mat. Pum ed Appl., 40 (1955), 61-67.
[Ku3]
Kuratowski K., Sur une method de metrisation complete, Fund. Math., 43 (1956), 114-138.
[LR]
Lasry J.M., Robert J., Analyse non lineaire multivoque, Pub!. no 7611, Centre Rech. Math. Decision (Ceremade), Universite de Paris IX (Dauphine), 71-190.
509 [LS]
LaSalle J.P., Stability theory and invariance principles, Dinamical systems an International Symposium, 1 (1976), 65-74, Academic Press, N.Y..
[Ma]
Markushevic A.I., Analytic functions theory, 2, Nauka, Moscow, 1968 (Russian).
[Mar]
Marcus L., Asymptoticaly autonomuous differential systems, Contributions to the Theory of Nonlinear Oscilations, III, 17-29, Annals of Math. Stud., N36, Princeton Univ. Press, N.Y., 1956.
[Me]
Mathematical Encyclopaedia, 5, SE, Moscow, 1985 (Russian).
[MS]
Miller R.K, Sell G.R, Topological dinamics and its relation to integral equations and nonautonomuous systems, Dinamical systems an International Symposium, 1 (1976), 223-249, Academic Press, N.Y..
[RHL]
Rouche N., Habets P., Laloy M., Stability theory by Liapounov's direct method, Springer-Verlag, New York-Heidelberg-Berlin, 1977.
[Se1]
Sell G.R, Nonautonomuous differential equations and topological dynamics I and II, Trans. Amer. Math. Soc., 127 (1967), 241-283.
[Se2]
Sell G.R., Lectures on Topological Dynamics and Differential Equations, Van Nostrand-Reinhold, London, 1971.
[Sc1]
Scorza-Dragoni G., Un teorema sulle funzioni continue rispetto ad una e misurabili rispetto ad un'altra variabile, Rend. Semin. mat. Univ. Padova, 17 (1948), 102106.
[Sc2]
Scorza-Dragoni G., Una applicatione della quasi continuitta semiregolare delle funzioni misurabili rispetto ad una e continue rispetto ad un'altra variabile, Atti Acad. Naz. Lincei, Ser.8, classe jis.,mat., 12, N1 (1952), 55-61.
[Yo]
Yorke J., Spaces of solutions, Mathematicals Systems Theory and Economics. II. Lect. Notes in Operst. Res. and Math. Econom. 12 (1969), 383-403.
[Za]
Zaremba S.K., Sur certaines familles de courbes en relation avec la theorie des equations differentielles, Rocznik Polskedo tow. matemat., XV (1936), 83-100.
INDEX
T2 -space, 18 Z-neighborhood, 267 Z-retraction, 317 IP-w-thin set, 232 IP-thin set, 232 c:-neighborhood, 15 c:-neighborhood of a set, 47 cp-w-thin set, 232 cp-thin set, 232 e-lifting, 63 e-Ioop, 70 e-homotopic mappings, 63 e- homotopy, 63 absolutely continuous function, 118 adherent point, 13 admissible change of variables, 292 admissible control, 353 Alexander lemma, 92 anti-discrete topology, 9 approximate derivative, 139 arcwise connected space, 61 argument, 2 Arzela theorem, 88 asymptoticaly stable point, 458 at most countable set, 7 at most countable set with respect to a space, 235 autonomous equation, 441 autonomous space, 441 Baire theorem, 26 ball, 15 Banach space, 72 base, 11 base of neighborhoods of a set, 29
bijective mapping, 5 boundary of a set, 13 bounded metric space, 47 Brouwer Fixed Point Theorem, 70 Cantor perfect set, 25 Cantor stairs, 113 Caratheodory conditions, 255 cardinality, 6 Cartesian product, 3 Cauchy problem for a differential inclusion (equation), 199 Cauchy problem for a space, 202 Cauchy sequence, 25 characteristic polynomial, 335 characteristic vector, 501 closed ball, 15 closed convex hull, 51 closed cover, 23 closed curve, 61 closed path, 61, 62 closed set, 13 closure, 13 compact open topology, 73 compact space, 23 compactum, 23 complement, 3 complete space, 25 component, 59 conic space, 457 connected component, 59 connected space, 56 constant mapping, 5 contingent, 268 continuous mapping, 30
511 continuous mapping (at a point), 27 continuum, 58 contractive mapping, 31 convergence of a sequence of points to a set, 399 convergence of series, 73 convergence of space sequences, 213 convergent generalized sequence, 22 convex combinations, 52 convex hull, 51 convex set, 50 countable set, 7 cover, 23 curve, 61 Davy conditions, 255 de Morgan formulae, 4 decreasing function, 108 definite integral, 128 degree of a mapping, 66 Denjoy integrable function, 140 dense subset, 26 density point, 138 derivative of a function with respect to a set, 116 diameter of a set, 39 diameter of a set with respect to a space, 442 difference of sets, 3 differential equation with multivalued right hand side, 193 differential inclusions, 192 directed set, 9 disconnected space, 56 discrete topology, 9 distance, 14 distance between subsets, 17 domain, 4 domain (of definition) , 2
domain of definition, 4 Egorov theorem, 161 element of a set, 1 embedding, 43 empty set, 2 end of a path, 60 end of a simple arc, 62 endpoints, 50 endpoints of a path, 60 equi-absolutely continuous set, 358 equi-absolutely continuous sequence of functions, 183 equicontinuous set of mappings, 87 equivalent sets, 6 Euclidean distance, 14 everywhere dense subset, 26 extendable mapping, 313 factor, 3, 38 family, 2 Fatou lemma, 137 filter, 9 finite intersection property, 24 finite set with respect to a space, 235 first approximation, 457 fixed point, 31, 69 function, 2 fundamental sequence, 25 fundamental system of solutions, 331 Fundamental theorem of algebra,
68 general solution, 264 generalized sequence, 17 graph, 5 Hahn-Banach Theorem, 166 Hausdorff distance, 80 Hausdorff metric, 80
512 Hausdorff space, 18 Hilbert cube, 45 homeomorphic spaces, 42 homogeneous linear equation, 328, 334 identity mapping, 5 image, 2 improper subset, 2 increasing function, 108 indefinite integral, 128, 141 index, 3 index of a point with respect to a loop, 484 index of a point with respect to a space, 485 index of a space with respect to a loop, 484 index set, 2 ind uced topology, 14 initial point of a path, 60 initial point of a simple arc, 62 injective mapping, 5 integral, 128 interior, 12 interior point, 12 interval of a circle, 470 inverse mapping, 5 isolated point, 21 Jordan form, 333 Lebesgue integrable function, 128 Lebesgue lemma, 490 Lebesgue theorem, 136 left Z-neighborhood, 267 left contingent, 268 left increasing point, 109 Lienard transformation, 407 limit of a generalized sequence, 18 limit ordinal number, 36 limit point of a generalized sequence, 17
limit point of a set, 17 linear functional, 50 Liouville's formula, 332 Lipschitz constant, 30 Lipschitz condition, 30 locally arcwise connected space, 61 locally compact space, 49 locally integrable function, 129 loop, 61, 62 loop bounding a region, 62 Luzin theorem, 161 majorant, 271 mapping, 2 mapping onto ... , 5 measurable function, 104 measurable multi-valued mapping, 152 measurable set, 99 measure of a set, 99 measure of a set with respect to a space, 232 metric, 14 Minkowski functional, 169 monotone function, 109 nearest point, 52 negative change of variables, 292 neighborhood, 12 Nemytski plane, 11 non-characteristic vector, 501 non-decreasing function, 108 non-homogeneous linear equation, 328, 334 non-increasing function, 108 non-isolated point, 21 nontrivial connected set, 58 norm, 50 normal space, 48 normed space, 50 nowhere dense subset, 26
513 one to one mapping, 5 open ball, 15 open cover, 23 open-closed set, 56 ordinal number, 35 ordinary differential equation, 191 outer measure, 96 partial order, 33 partial ordering, 33 path, 60 path connecting points, 60 Peano condition, 266 periodic space, 452 Picard condition, 267 Poincare-Bendixson theorem, 498 positive change of variables, 292 precedes, 33 preimage,2 primitive, 128, 141 product of a family of sets, 3 projection, 38 proper subset, 2 region, 61 regular space, 28 relatively compact subset, 23 relatively weakly compact sequence, 180 restriction, 5 retraction, 317 right Z-neighborhood, 267 right contingent, 268 right increasing point, 109 Schauder Fixed Point Theorem, 70 Scorza-Dragoni theorem, 160 Scorza-Dragoni condition, 280 segment, 50 segment of a circle, 63 self-intersection, 61 sequence, 7
series, 73 set, 1 set free of stationary points, 483 set of measure zero, 103 simple arc, 62 simple closed arc, 62 simple closed path, 62 simple loop, 61, 62 simple path, 61 singleton, 8 slitting set, 61 solution of a control problem, 353 solution of a control problem with restrictions, 353 solution of a differential inclusion (equation ), 193 solution space, 193 space measurable sets, 179 space of compact subsets, 79 stationary point of a solution space, 443 strongly monotone function, 109 sub-base of neighborhoods of a set, 29 subbase, 12 subcover, 23 subproduct, 38 subsequence, 7 subset, 2 subset of full measure, 104 subspace of a topological space, 14 subspace of a metric space, 17 succeeds, 33 sufficient me-approximation, 274 sufficient mek-approximation, 274 sum of series, 73 surjective mapping, 5 system of neighborhoods, 10 topological limit, 22 total length, 95 transfinite, 35
514 triangle inequality, 14 trivial connected set, 58 Tychonoff neighborhood, 37 Tychonoff topology, 37 Tychonov theorem, 93 uniformly continuous family of functions, 88 uniformly convergent series, 73 upper semicontinuous mapping, 30 upper semicontinuous mapping (at a point), 27 value of a function (mapping), 2 variation of constants, 334 Vietoris neighborhood, 77 Viet oris topology, 77 weak topology of a normed space, 168 weak topology on L *, 165 weakly convergent sequences, 171 Wronskian, 331 Zermelo theorem, 34 Zorgenfrei line, 10 Zorn lemma, 33
NOTATION
(+), 292, 313 (-), 292, 313
(X, p), 14 (v), 368 (c), 197 (e), 203 (1), 309 (k), 247 (l), 309 (n'), 198 (n), 196 (p), 195 (q), 194 (r), 309 (8), 195 (u), 203 (x, e) - inner (scalar) product, 50 A(V),441 A 3 8, 2 A ;i 8, 2 Ag B, 2 A ~ B, 2 A*(V), 441 B(V), 444 B(X, Y), 71 BC(X, Y), 71 B P- A, 2 B ;;2 A, 2 B*(V), 444 C(X, Y), 72 Cv(X, Y), 83 Cvc(X, Y), 83 D(F), 193 D(F,M),193 DN - Cantor perfect set, 46 F",198
Ft, 198 L I , 173 L I ([a,b]),173 Li([a, b]), 399 L oo , 174 MY, 198 MI (h, c), 233 M 2 (h, c), 234 Mt, 198 O(K,U),73 O(x, E), 15 O(UI , ... , Uk), 77 O,(M,E),47 O,(x, E), 15 OEX, 15 pa(u), 452 P:(U), 452 Q**G(U), 271 Q~*G(U), 271
Q**(U), 271 Q~*(U), 271 Q*G(U), 271 Q*(U), 198 Qt(U),211 Qd(
Z+,304 Z-,304
516
Z+,204 Z-+, 205, 306 Z-,204 zne(+J, 308 zne(-J, 308 zne, 308 ZM,195
[MJ, 13 [MJx, 13 8M, 13 Gr(f),5 Im(f),5 n+(z), 446 n-(z), 446 Re x - real part of x, 461 Stat(Z), 483 c(M), 51 n'Y,3 n{XI , ... ,Xn }, 3 n{Xa: a E A}, 3 ni'=lXi ,3 naEAXa, 3 cc(M),51 cont(Z, x), 268 contr(Z, x), 268 contl(Z, x), 268 U'Y, 3 U{XI , ... ,Xn }, 3 U{Xa: a E A}, 3 Ui'=IX i ,3 UaEAXa, 3 0,2 expc X, 78 expc X, 79 (M), 12 (M)x' 12 (Z, Zoo), 444 lim top, 22 lim top inf, 19 lim top sup, 19 J.L*(M), 95 J.L~(M),
wo,36
232
wI,36 w~(M), 232 w,AM),232 7r(f) = A, 4 7rb, 84 7re , 84 -<, 33 I1'Y,3 I1{X I , ... ,Xn },3 I1{Xa: a E A}, 3 I1~=1 X k , 3 I1Z=1 X k , 3 I1aEA X a, 3 p(A, B), 17 >-, 33 r Z , 267 <j;, 291
db),95
j(M),2 j-I(M), 2 j(Z, 1),484 j(Z, j, c), 483 jz(a), 485 l1f' 84 s(U), 213 sEA,2 s ~ A, 2 x = lim.., x a , 18 Xa --t x, 18 y' = (t, y), 192 y' E F(t, y), 192 e - set of complex numbers, complex plane, 68 e - set of complex numbers, complex plane, 192 en, 192 N - set of positive integers, 7 Q - set of rational numbers, 8 lR - real line, 9 lRn n-dimensional Euclidean space, 4 Z - set of integers, 8 (n*), 267
Other Mathematics and Its Applications titles of interest:
I.H. Dimovski: Convolutional Calculus. 1990, 208 pp.
ISBN 0-7923-0623-6
Y.M. Svirezhev and V.P. Pasekov: Fundamentals of Mathematical Evolutionary Genetics. 1990,384 pp. ISBN 90-277-2772-4 S. Levendorskii: Asymptotic Distribution of Eigenvalues of Differential Operators. 1991, 297 pp. ISBN 0-7923-0539-6 V.G. Makhankov: Soliton Phenomenology. 1990,461 pp.
ISBN 90-277-2830-5
I. Cioranescu: Geometry of Banach Spaces. Duality Mappings and Nonlinear Problems. 1990,274 pp. ISBN 0-7923-0910-3 B.1. Sendov: Hausdorff Approximation. 1990,384 pp.
ISBN 0-7923-0901-4
A.B. Venkov: Spectral Theory of Automorphic Functions and Its Applications. 1991,280 pp. ISBN 0-7923-0487-X V.I. Arnold: Singularities of Caustics and Wave Fronts. 1990,274 pp. ISBN 0-7923-1038-1 A.A. Pankov: Bounded and Almost Periodic Solutions of Nonlinear Operator Differential Equations. 1990, 232 pp. ISBN 0-7923-0585-X A.S. Davydov: Solitons in Molecular Systems. Second Edition. 1991, 428 pp. ISBN 0-7923-1029-2 B.M. Levitan and I.S. Sargsjan: Sturm-Liouville and Dirac Operators. 1991, 362 pp. ISBN 0-7923-0992-8 V.1. Gorbachuk and M.L. Gorbachuk: Boundary Value Problems for Operator Differential Equations. 1991, 376 pp. ISBN 0-7923-0381-4 Y.S. Samoilenko: Spectral Theory of Families of Self-Adjoint Operators. 1991, 309 pp. ISBN 0-7923-0703-8 B.1. Golubov A.V. Efimov and V.A. Scvortsov: Walsh Series and Transforms. 1991,382 pp. ISBN 0-7923-1100-0 V. Laksmikantham, V.M. Matrosov and S. Sivasundaram: Vector Lyapunov Functions and Stability Analysis of Nonlinear Systems. 1991, 250 pp. ISBN 0-7923-1152-3 F.A. Berezin and M.A. Shubin: The Schrodinger Equation. 1991,556 pp. ISBN 0-7923-1218-X D.S. Mitrinovic, J.E. Pecaric and A.M. Fink: Inequalities Involving Functions and their Integrals and Derivatives. 1991, 588 pp. ISBN 0-7923-1330-5 Julii A. Dubinskii: Analytic Pseudo-Differential Operators and their Applications. 1991,252 pp. ISBN 0-7923-1296-1 V.I. Fabrikant: Mixed Boundary Value Problems in Potential Theory and their ISBN 0-7923-1157-4 Applications. 1991,452 pp.
Other Mathematics and Its Applications titles of interest:
A.M. Samoilenko: Elements of the Mathematical Theory of Multi-Frequency Oscillations. 1991, 314 pp. ISBN 0-7923-1438-7 Yu.L. Dalecky and S.V. Fomin: Measures and Differential Equations in InfiniteDimensional Space. 1991,338 pp. ISBN 0-7923-1517-0 W. Mlak: Hilbert Space and Operator Theory. 1991, 296 pp. ISBN 0-7923-1042-X N.Ja. Vilenkin and A.V. Klimyk: Representation of Lie Groups and Special Functions. Volume 1: Simplest Lie Groups, Special Functions, and Integral Transforms. 1991,608 pp. ISBN 0-7923-1466-2 N.Ja. Vilenkin and A.V. Klimyk: Representation of Lie Groups and Special Functions. Volume 2: Class I Representations, Special Functions, and Integral Transforms. 1992, 630 pp. ISBN 0-7923-1492-1 N.la. Vilenkin and A.V. Klimyk: Representation of Lie Groups and Special Functions. Volume 3: Classical and Quantum Groups and Special Functions. 1992, 650 pp. ISBN 0-7923-1493-X (Set ISBN for Vols. 1,2 and 3: 0-7923-1494-8) K. Gopalsamy: Stability and Oscillations in Delay Differential Equations of Population Dynamics. 1992,502 pp. ISBN 0-7923-1594-4 N.M. Korobov: Exponential Sums and their Applications. 1992,210 pp. ISBN 0-7923-1647-9 Chuang-Gan Hu and Chung-Chun Yang: Vector-Valued Functions and their Applications. 1991, 172 pp. ISBN 0-7923-1605-3 Z. Szmydt and B. Ziemian: The Mellin Transformation and Fuchsian Type Partial Differential Equations. 1992, 224 pp. ISBN 0-7923-1683-5 L.I. Ronkin: Functions of Completely Regular Growth. 1992, 394 pp. ISBN 0-7923-1677-0 R. Delanghe, F. Sommen and V. Soucek: Clifford Algebra and Spinor-valued Functions. A Function Theory of the Dirac Operator. 1992, 486 pp. ISBN 0-7923-0229-X A. Tempelman: Ergodic Theorems for Group Actions. 1992, 400 pp. ISBN 0-7923-1717-3 D. Bainov and P. Simenov: Integral Inequalities and Applications. 1992, 426 pp. ISBN 0-7923-1714-9 I. Imai: Applied Hyperfunction Theory. 1992, 460 pp.
ISBN 0-7923-1507-3
Yu.1. Neimark and P.S. Landa: Stochastic and Chaotic Oscillations. 1992,502 pp. ISBN 0-7923-1530-8 H.M. Srivastava and R.G. Buschman: Theory and Applications of Convolution Integral Equations. 1992,240 pp. ISBN 0-7923-1891-9
Other Mathematics and Its Applications titles of interest:
A. van der Burgh and J. Simonis (eds.): Topics in Engineering Mathematics. 1992, 266 pp. ISBN 0-7923-2005-3 F. Neuman: Global Properties of Linear Ordinary Differential Equations. 1992, 320 pp. ISBN 0-7923-1269-4 A. Dvurecenskij: Gleason's Theorem and its Applications. 1992, 334 pp. ISBN 0-7923-1990-7 D.S. Mitrinovic, J.E. Pecaric and A.M. Fink: Classical and New Inequalities in Analysis. 1992, 740 pp. ISBN 0-7923-2064-6 H.M. Hapaev: Averaging in Stability Theory. 1992,280 pp.
ISBN 0-7923-1581-2
S. Gindinkin and L.R. Volevich: The Method of Newton's Polyhedron in the Theory ofPDE's. 1992,276 pp. ISBN 0-7923-2037-9 Yu.A. Mitropolsky, A.M. Samoilenko and 0.1. Martinyuk: Systems of Evolution Equations with Periodic and Quasiperiodic Coefficients. 1992, 280 pp. ISBN 0-7923-2054-9 LT. Kiguradze and T.A. Chanturia: Asymptotic Properties of Solutions of Nonautonomous Ordinary Differential Equations. 1992, 332 pp. ISBN 0-7923-2059-X V.L. Kocic and G. Ladas: Global Behavior of Nonlinear Difference Equations of Higher Order with Applications. 1993, 228 pp. ISBN 0-7923-2286-X S. Levendorskii: Degenerate Elliptic Equations. 1993,445 pp. ISBN 0-7923-2305-X D. Mitrinovic and J.D. Keckic: The Cauchy Method of Residues, Volume 2. Theory and Applications. 1993, 202 pp. ISBN 0-7923-2311-8 R.P. Agarwal and P.J.Y Wong: Error Inequalities in Polynomial Interpolation and Their Applications. 1993, 376 pp. ISBN 0-7923-2337-8 A.G. Butkovskiy and L.M. Pustyl'nikov (eds.): Characteristics of DistributedParameter Systems. 1993,386 pp. ISBN 0-7923-2499-4 B. Stemin and V. Shatalov: Differential Equations on Complex Manifolds. 1994, 504 pp. ISBN 0-7923-2710-1 S.B. Yakubovich and Y.F. Luchko: The Hypergeometric Approach to Integral Transforms and Convolutions. 1994, 324 pp. ISBN 0-7923-2856-6 C. Gu, X. Ding and C.-c. Yang: Partial Differential Equations in China. 1994, 181 pp. ISBN 0-7923-2857-4
V.G. Kravchenko and G.S. Litvinchuk: Introduction to the Theory of Singular ISBN 0-7923-2864-7 Integral Operators with Shift. 1994, 288 pp. A. Cuyt (ed.): Nonlinear Numerical Methods and Rational Approximation II. 1994, 446 pp. ISBN 0-7923-2967-8
Other Mathematics and Its Applications titles of interest:
G. Gaeta: Nonlinear Symmetries and Nonlinear Equations. 1994, 258 pp. ISBN 0-7923-3048-X V.A. Vassiliev: Ramified Integrals, Singularities and Lacunas. 1995,289 pp. ISBN 0-7923-3193-1 NJa. Vilenkin and A.V. Klimyk: Representation of Lie Groups and Special Functions. Recent Advances. 1995,497 pp. ISBN 0-7923-3210-5 Yu. A. Mitropolsky and A.K. Lopatin: Nonlinear Mechanics, Groups and Symmetry. 1995, 388 pp. ISBN 0-7923-3339-X R.P. Agarwal and P.Y.H. Pang: Opiallnequalities with Applications in Differential and Difference Equations. 1995, 393 pp. ISBN 0-7923-3365-9 A.G. Kusraev and S.S. Kutateladze: Subdifferentials: Theory and Applications. 1995, 408 pp. ISBN 0-7923-3389-6 M. Cheng, D.-G. Deng, S. Gong and c.-c. Yang (eds.): Harmonic Analysis in China. 1995, 318 pp. ISBN 0-7923-3566-X M.S. Livsic, N. Kravitsky, A.S. Markus and V. Vinnikov: Theory of Commuting Nonselfadjoint Operators. 1995, 314 pp. ISBN 0-7923-3588-0 A.1. Stepanets: Classification and Approximation of Periodic Functions. 1995, 360 pp. ISBN 0-7923-3603-8 c.-G. Ambrozie and F.-H. Vasilescu: Banach Space Complexes. 1995,205 pp. ISBN 0-7923-3630-5 E. Pap: Null-Additive Set Functions. 1995, 312 pp.
ISBN 0-7923-3658-5
CJ. Colboum and E.S. Mahmoodian (eds.): Combinatorics Advances. 1995, 338 pp. ISBN 0-7923-3574-0 V.G. Danilov, V.P. Maslov and K.A. Volosov: Mathematical Modelling of Heat and Mass Transfer Processes. 1995, 330 pp. ISBN 0-7923-3789-1 A. LaurinCikas: Limit Theorems for the Riemann Zeta-Function. 1996, 312 pp. ISBN 0-7923-3824-3 A. Kuzhel: Characteristic Functions and Models of NonselfAdjoint Operators. 1996, 283 pp. ISBN 0-7923-3879-0 G.A. Leonov, I.M. Burkin and A.1. Shepeljavyi: Frequency Methods in Oscillation Theory. 1996,415 pp. ISBN 0-7923-3896-0 B. Li, S. Wang, S. Yan and c.-c. Yang (eds.): Functional Analysis in China. 1996, 390 pp. ISBN 0-7923-3880-4 P.S. Landa: Nonlinear Oscillations and Waves in Dynamical Systems. 1996, 554 pp. ISBN 0-7923-3931-2
Other Mathematics and Its Applications titles of interest:
AJ. Jerri: Linear Difference Equations with Discrete Transform Methods. 1996,
462 pp.
ISBN 0-7923-3940-1
I. Novikov and E. Semenov: Haar Series and Linear Operators. 1997,234 pp.
ISBN 0-7923-4006-X L. Zhizhiashvili: Trigonometric Fourier Series and Their Conjugates. 1996, 312 pp. ISBN 0-7923-4088-4 R.G. Buschman: Integral Transformation, Operational Calculus, and Generalized Functions. 1996,246 pp. ISBN 0-7923-4183-X V. Lakshmikantham, S. Sivasundaram and B. Kaymakcalan: Dynamic Systems on Measure Chains. 1996,296 pp. ISBN 0-7923-4116-3 D. Guo, V. Lakshmikantham and X. Liu: Nonlinear Integral Equations in Abstract Spaces. 1996,350 pp. ISBN 0-7923-4144-9 Y. Roitberg: Elliptic Boundary Value Problems in the Spaces of Distributions. 1996, 427 pp. ISBN 0-7923-4303-4 Y. Komatu: Distortion Theorems in Relation to Linear Integral Operators. 1996, 313 pp. ISBN 0-7923-4304-2 A.G. Chentsov: Asymptotic Attainability. 1997,336 pp.
ISBN 0-7923-4302-6
S.T. Zavalishchin and A.N. Sesekin: Dynamic Impulse Systems. Theory and Applications. 1997,268 pp. ISBN 0-7923-4394-8 U. Elias: Oscillation Theory of Two-Term Differential Equations. 1997,226 pp. ISBN 0-7923-4447-2 D. O'Regan: Existence Theory for Nonlinear Ordinary Differential Equations. 1997, 204 pp. ISBN 0-7923-4511-8 Yu. Mitropolskii, G. Khoma and M. Gromyak: Asymptotic Methods for Investigating Quasiwave Equations of Hyperbolic Type. 1997,418 pp. ISBN 0-7923-4529-0 R.P. Agarwal and P.J.Y. Wong: Advanced Topics in Difference Equations. 1997, 518 pp. ISBN 0-7923-4521-5 N.N. Tarkhanov: The Analysis of Solutions of Elliptic Equations. 1997, 406 pp. ISBN 0-7923-4531-2 B. Rieean and T. Neubrunn: Integral, Measure, and Ordering. 1997, 376 pp. ISBN 0-7923-4566-5 N.L. Gol'dman: Inverse Stefan Problems. 1997,258 pp.
ISBN 0-7923-4588-6
S. Singh, B. Watson and P. Srivastava: Fixed Point Theory and Best ApproximaISBN 0-7923-4758-7 tion: The KKM-map Principle. 1997, 230 pp. A. Pankov: G-Convergence and Homogenization of Nonlinear Partial Differential ISBN 0-7923-4720-X Operators. 1997,263 pp.
Other Mathematics and Its Applications titles of interest:
S. Hu and N.S. Papageorgiou: Handbook of Multivalued Analysis. Volume I: Theory. 1997, 980 pp. ISBN 0-7923-4682-3 (Set of 2 volumes: 0-7923-4683-1) L.A. Sakhnovich: Interpolation Theory and Its Applications. 1997, 216 pp. ISBN 0-7923-4830-0 G. V. Milovanovic: Recent Progress in Inequalities. 1998, 531 pp. ISBN 0-7923-4845-1 V.V. Filippov: Basic Topological Structures of Ordinary Differential Equations. 1998,530 pp. ISBN 0-7293-4951-2 S. Gong: Convex and Starlike Mappings in Several Complex Variables. 1998, 208 pp. ISBN 0-7923-4964-4