G1CMIN Measure and Integration

Similar documents
Logical Connectives and Quantifiers

Chapter 4. Measure Theory. 1. Measure Spaces

Analysis Finite and Infinite Sets The Real Numbers The Cantor Set

Theorems. Theorem 1.11: Greatest-Lower-Bound Property. Theorem 1.20: The Archimedean property of. Theorem 1.21: -th Root of Real Numbers

MAT 570 REAL ANALYSIS LECTURE NOTES. Contents. 1. Sets Functions Countability Axiom of choice Equivalence relations 9

In N we can do addition, but in order to do subtraction we need to extend N to the integers

5 Set Operations, Functions, and Counting

In N we can do addition, but in order to do subtraction we need to extend N to the integers

A LITTLE REAL ANALYSIS AND TOPOLOGY

Lebesgue measure and integration

MATH31011/MATH41011/MATH61011: FOURIER ANALYSIS AND LEBESGUE INTEGRATION. Chapter 2: Countability and Cantor Sets

1/12/05: sec 3.1 and my article: How good is the Lebesgue measure?, Math. Intelligencer 11(2) (1989),

Chapter 1 The Real Numbers

Some Background Material

Real Analysis Math 131AH Rudin, Chapter #1. Dominique Abdi

REAL VARIABLES: PROBLEM SET 1. = x limsup E k

Week 2: Sequences and Series

Analysis I. Classroom Notes. H.-D. Alber

MATHS 730 FC Lecture Notes March 5, Introduction

Compendium and Solutions to exercises TMA4225 Foundation of analysis

Chapter 1. Measure Spaces. 1.1 Algebras and σ algebras of sets Notation and preliminaries

MATH 131A: REAL ANALYSIS (BIG IDEAS)

We are going to discuss what it means for a sequence to converge in three stages: First, we define what it means for a sequence to converge to zero

Measure and integration

MATH 5616H INTRODUCTION TO ANALYSIS II SAMPLE FINAL EXAM: SOLUTIONS

Lecture 3: Probability Measures - 2

MATH41011/MATH61011: FOURIER SERIES AND LEBESGUE INTEGRATION. Extra Reading Material for Level 4 and Level 6

REVIEW OF ESSENTIAL MATH 346 TOPICS

Chapter 1. Sets and Mappings

MATH5011 Real Analysis I. Exercise 1 Suggested Solution

Problem set 1, Real Analysis I, Spring, 2015.

MATH 102 INTRODUCTION TO MATHEMATICAL ANALYSIS. 1. Some Fundamentals

Lebesgue Measure. Dung Le 1

Measures. Chapter Some prerequisites. 1.2 Introduction

Introduction to Proofs in Analysis. updated December 5, By Edoh Y. Amiran Following the outline of notes by Donald Chalice INTRODUCTION

Indeed, if we want m to be compatible with taking limits, it should be countably additive, meaning that ( )

Lebesgue Measure on R n

FUNDAMENTALS OF REAL ANALYSIS by. II.1. Prelude. Recall that the Riemann integral of a real-valued function f on an interval [a, b] is defined as

Introduction to Real Analysis

van Rooij, Schikhof: A Second Course on Real Functions

Measures. 1 Introduction. These preliminary lecture notes are partly based on textbooks by Athreya and Lahiri, Capinski and Kopp, and Folland.

Measure Theory and Lebesgue Integration. Joshua H. Lifton

Quick Tour of the Topology of R. Steven Hurder, Dave Marker, & John Wood 1

Copyright c 2007 Jason Underdown Some rights reserved. statement. sentential connectives. negation. conjunction. disjunction

Mathematics Course 111: Algebra I Part I: Algebraic Structures, Sets and Permutations

Mathematical Methods for Physics and Engineering

Notes for Math 290 using Introduction to Mathematical Proofs by Charles E. Roberts, Jr.

Topology. Xiaolong Han. Department of Mathematics, California State University, Northridge, CA 91330, USA address:

MATS113 ADVANCED MEASURE THEORY SPRING 2016

1.4 Outer measures 10 CHAPTER 1. MEASURE

Real Analysis Chapter 1 Solutions Jonathan Conder

Undergraduate Notes in Mathematics. Arkansas Tech University Department of Mathematics

MATH NEW HOMEWORK AND SOLUTIONS TO PREVIOUS HOMEWORKS AND EXAMS

STAT 7032 Probability Spring Wlodek Bryc

Chapter 3 Continuous Functions

Solutions Manual for: Understanding Analysis, Second Edition. Stephen Abbott Middlebury College

µ (X) := inf l(i k ) where X k=1 I k, I k an open interval Notice that is a map from subsets of R to non-negative number together with infinity

Solutions to Tutorial 1 (Week 2)

MATH 409 Advanced Calculus I Lecture 10: Continuity. Properties of continuous functions.

Contents Ordered Fields... 2 Ordered sets and fields... 2 Construction of the Reals 1: Dedekind Cuts... 2 Metric Spaces... 3

Integration on Measure Spaces

JOURNAL OF INQUIRY-BASED LEARNING IN MATHEMATICS

Part V. 17 Introduction: What are measures and why measurable sets. Lebesgue Integration Theory

Measures and Measure Spaces

1.1. MEASURES AND INTEGRALS

Principle of Mathematical Induction

(U) =, if 0 U, 1 U, (U) = X, if 0 U, and 1 U. (U) = E, if 0 U, but 1 U. (U) = X \ E if 0 U, but 1 U. n=1 A n, then A M.

Sequences. Chapter 3. n + 1 3n + 2 sin n n. 3. lim (ln(n + 1) ln n) 1. lim. 2. lim. 4. lim (1 + n)1/n. Answers: 1. 1/3; 2. 0; 3. 0; 4. 1.

Measure and Integration: Concepts, Examples and Exercises. INDER K. RANA Indian Institute of Technology Bombay India

REAL AND COMPLEX ANALYSIS

Week Some Warm-up Questions

The Real Number System

van Rooij, Schikhof: A Second Course on Real Functions

II - REAL ANALYSIS. This property gives us a way to extend the notion of content to finite unions of rectangles: we define

R N Completeness and Compactness 1

THEOREMS, ETC., FOR MATH 515

3 COUNTABILITY AND CONNECTEDNESS AXIOMS

Chapter 1. Sets and Numbers

3 Measurable Functions

Daniel Akech Thiong Math 501: Real Analysis Homework Problems with Solutions Fall Problem. 1 Give an example of a mapping f : X Y such that

Contribution of Problems

Spring 2014 Advanced Probability Overview. Lecture Notes Set 1: Course Overview, σ-fields, and Measures

Economics 204 Fall 2011 Problem Set 1 Suggested Solutions

Extension of continuous functions in digital spaces with the Khalimsky topology

Sets, Structures, Numbers

3 (Due ). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure?

Axioms for Set Theory

Section 2.5 : The Completeness Axiom in R

Elementary Analysis Math 140D Fall 2007

ADVANCED CALCULUS - MTH433 LECTURE 4 - FINITE AND INFINITE SETS

Partial Solutions to Folland s Real Analysis: Part I

DR.RUPNATHJI( DR.RUPAK NATH )

+ = x. ± if x > 0. if x < 0, x

Structure of R. Chapter Algebraic and Order Properties of R

ABOUT THE CLASS AND NOTES ON SET THEORY

One-to-one functions and onto functions

Lebesgue Integration: A non-rigorous introduction. What is wrong with Riemann integration?

Math 117: Topology of the Real Numbers

MATH FINAL EXAM REVIEW HINTS

M17 MAT25-21 HOMEWORK 6

Transcription:

G1CMIN Measure and Integration 2003-4 Prof. J.K. Langley May 13, 2004 1 Introduction Books: W. Rudin, Real and Complex Analysis ; H.L. Royden, Real Analysis (QA331). Lecturer: Prof. J.K. Langley (jkl@maths, room C6c, 95 14964). Lectures: Wed 11 C29; Th 1 C4; Th 5 C29. There will be NO lectures in the week commencing Feb 9th. Office hours: none specified but my door is open most of the time and you re free to consult me. Assessment: 2.5 hour written examination. 5 questions, best 4 count. PLAS NOT: G1CMIN is a traditional pure mathematics theory module along the familiar definition, theorem, proof structure; in particular there isn t much scope for calculations and the module is more like G12RAN or G13MTS than G12CAN. Contents: review of real analysis, Lebesgue measure, Lebesgue integration. Outline: there are two main themes to the module. The idea of measure is concerned with the size of sets. The first distinction we meet between sets is usually between finite and infinite sets. We then refine the idea of an infinite set to distinguish between the countable and the uncountable (we ll review this concept). When we discuss the Lebesgue measure of a subset of R, it will give us an indication of how much of the line is filled up by our set. Thus we will be able to distinguish big uncountable sets from smaller ones. The second main theme involves integration. G12RAN introduces the Riemann integral, which has advantages in that it is relatively easy to define, and displays well the link between integration and differentiation. Its main drawbacks are: (i) the class of Riemann integrable functions is too small; (ii) it has technical problems, particularly with regard to whether lim ( n f n ) = Also, Riemann s integral is difficult to generalize to other settings. ( lim n f n ). (1) 1

Lebesgue measure gives a means for comparing the size of sets and leads to Lebesgue integration, which is used widely in pure maths, probability, mathematical physics, PDs etc. This module will go as far as the main theorems concerning when (1) holds. More advanced topics will only be covered if time permits. Aims and Learning Outcomes: Aims: to teach the elements of measure theory and Lebesgue integration. Learning Outcomes: a successful student will: 1. be able to state, and apply in the investigation of examples, the principal theorems as treated; 2. be able to prove simple propositions concerning sets, measure spaces, Lebesgue measure and integrable functions. Coursework: is not part of the assessment for G1CMIN, but the questions should be helpful practice for the exam. There will be a short assignment each week, for handing in the week after. For anyone taking this module for G13S1 (Supplementary Maths), the assessment will consist of a separate coursework assignment handed out by the end of Week 6 of the semester. It will be due for handing in on the last day of the Spring term and will be based on material from the first half of the module. You need only attend enough lectures to cover this material. Web notes: you can find printed notes at www.maths.nott.ac.uk/personal/jkl/min03.pdf The lectures will, however, cover all of the material (except for some proofs given in previous modules such as G1ALIM) so you may prefer to use the Web notes either not at all or just as backup. These notes may contain errors, omissions or obscure parts: these will be amended as and when I find them. 2

2 Sets 2.1 Countability This is an important idea when deciding how big infinite sets are compared to each other. We shall see that R is a bigger set than Q. We say that a set A is countable if either A is empty or there is a sequence (a n ), n = 1, 2, 3,..., which uses up A, by which we mean that each a n A and each member of A appears at least once in the sequence. This is the same as saying that there is a surjective (onto) function f : N A (via a n = f(n)). FACT 1: Any finite set is countable. Indeed, if A = {x 1,..., x N }, just put a n = x n if n N, and a n = x N if n > N. FACT 2: If B A and A is countable, then B is countable. If B is finite, this is obvious. If B is not finite, then nor is A, so take a sequence which uses up A, and delete all entries in the sequence which don t belong to B. We then get an infinite sequence which uses up B. FACT 3: Suppose that A is an infinite, countable set. Then there is a sequence (b n ), n = 1, 2,..., of members of A in which each member of of A appears exactly once. To see this, suppose that (a n ), n = 1, 2,... uses up A. Go through the list, deleting any entry which has previously occurred. So if a n = a j for some j < n, we delete a n. The resulting subsequence includes each member of A exactly once. We have thus arranged A into a sequence - first element, second element etc. - hence the name countable. FACT 4: Suppose that A 1, A 2, A 3,... are countably many countable sets. Then the union U = n=1 A n, which is the set of all x which each belong to at least one A n, is countable. Proof: delete any A j which are empty, and re-label the rest. Now suppose that the j th set A j is used up by the sequence (a j,n ), n = 1, 2,.... Write out these sequences as follows: a 1,1 a 1,2 a 1,3 a 1,4............... a 2,1 a 2,2 a 2,3 a 2,4............... a 3,1 a 3,2 a 3,3 a 3,4............... etc. Now the following sequence uses up all of U. We take a 1,1 a 1,2 a 2,1 a 1,3 a 2,2 a 3,1 a 1,4 a 2,3... FACT 5: The set of positive rational numbers is countable. The reason is that this set is the union of the sets A m = {p/m : p N}, each of which is countable. Similarly, the set of negative rational numbers is countable, and so is Q (the union of these two sets and {0}). FACT 6: If A and B are countable sets, then so is the Cartesian product A B, which is 3

the set of all ordered pairs (a, b), with a A and B B. Here ordered means that (a, b) (b, a) unless a = b. Fact 6 is obvious if A or B is empty. Otherwise, if (a n ) uses up A and (b n ) uses up B, then A B is the union of the sets C n = {(a n, b m ) : m = 1, 2, 3,...}, each of which is countable. FACT 7: the interval (0, 1) is not countable, and therefore nor are R, C. Proof: We prove the following stronger assertion. Consider the collection T of all real numbers x = 0 d 1 d 2 d 3 d 4...... in which each digit d j is either 4 or 5. Then T is uncountable. Suppose that the sequence (a n ), n = 1, 2,..., uses up T. Write out each a j as a decimal expansion a 1 = 0 b 1,1 b 1,2 b 1,3...... a 2 = 0 b 2,1 b 2,2 b 2,3...... a 3 = 0 b 3,1 b 3,2 b 3,3...... etc. Here each digit b j,k is 4 or 5. We make a new number x = 0 c 1 c 2 c 3 c 4... as follows. We look at b n,n. If b n,n = 4, we put c n = 5, while if b n,n = 5, we put c n = 4. Now x cannot belong to the list above, for if we had x = a m, then we d have c m = b m,m, which isn t true. xample Let S be the collection of all sequences a 1, a 2,... with each entry a positive integer. Then S is uncountable: take the subset of S for which each a j is 4 or 5 and use the above proof. 2.2 The real numbers R The key idea about R which we need is the existence of least upper bounds. Let be a non-empty subset of R. We say that a real number M is an upper bound for if x M for all x in, and is called bounded above. Among all upper bounds for there is one which is the least, called the sup or l.u.b. of. We adopt the convention that if is a subset of R which is not bounded above, then the sup of is +. For example sup((0, 1) Q) = 1 and sup N =. The greatest lower bound of (denoted glb or inf) is defined similarly: it is the greatest real 4

number which is less than or equal to every member of. If is not bounded below the glb is. 2.3 Real sequences Let (x n ), n = 1, 2,... be a sequence of real numbers. We say lim n x n = L R if to each positive real number ε corresponds an integer n 0 such that x n L < ε for all n n 0. We say lim n x n = + if to each positive real number M corresponds an integer n 0 such that x n > M for all n n 0. We say lim n x n = if lim n ( x n ) = +. Thus every negative real M has n 0 with x n < M for all n n 0. 2.4 The monotone sequence theorem If the real sequence (x n ) is non-decreasing (i.e. x n x n+1 ) for n N then x n tends to a limit (finite or + ). We recall the proof. Let s be the supremum of the set {x n : n N} = A. Suppose first that A is not bounded above, so that sup A = + by our convention. This means that if we are given some positive number M, then no matter how large M might be, we can find some member of the set A, say x n1, such that x n1 > M. But then, because the sequence is non-decreasing, we have x n > M for all n n 1, and this is precisely what we need in order to be able to say that lim n x n = +. Now suppose that A is bounded above, and let s be the sup. Suppose we are given some positive ε. Then we need to show that x n s < ε for all sufficiently large n. But we know that x n s for all n, so we just have to show that x n > s ε for all large enough n. This we do as follows. The number s ε is less than s and so is not an upper bound for A, and so there must be some n 2 such that x n2 > s ε. But then x n > s ε for all n n 2, and the proof is complete. 2.5 Lemma very real sequence (x n ), n = 1, 2,..., has a monotone subsequence. Proof: For each n consider the set n = {x m : m n}. We look at two cases. Suppose first that for every n the set n has a maximum element i.e. there is some m n such 5

that x p x m for all p n. To form our subsequence choose n 1 1 such that x n1 is the maximum element of 1. Then choose n 2 n 1 + 1 such that x n2 is the maximum element of n1 +1. Since n1 +1 1 we get x n2 x n1. Now take the maximum element of n2 +1: this will be x n3 for some n 3 > n 2 and we have x n3 x n2. Repeating this we get a non-increasing subsequence (x nk ). Suppose now that some N has no maximum element. Let n 1 = N and choose n 2 > n 1 such that x n2 > x n1. Since N has no maximum element, there is an element greater than all of x n1, x n1 +1,..., x n2, and this is x n3 for some n 3 > n 2. Carrying on in this way, we get a strictly increasing subsequence. 2.6 Corollary (Bolzano-Weierstrass theorem) very bounded real sequence has a bounded monotone subsequence and hence a convergent subsequence. 2.7 Nested intervals Let I k = [a k, b k ] be closed intervals in R such that I k+1 I k. Then a k a k+1 b k+1 b 1 and a 1 a k+1 b k+1 b k so a k converges, to A say, and b k converges, to B say. We have a k A B b k for all k, so [A, B] is contained in all of the I k. Thus the intersection of the I k is non-empty. The example I n = (0, 1/n) shows that open intervals do not in general have this property. 2.8 Open sets A subset U of R is called open if the following is true. To each x in U corresponds δ x > 0 such that (x δ x, x + δ x ) U. It is easy to check that if W t is open for each t in some set T then the set t T W t, which is the set of all y each belonging to at least one W t, is open. Also the intersection of finitely many open sets is open. 2.9 Open intervals By an open interval in R we mean any of the following: (a, b), (a, + ), (, b), R. All are open sets. A subset of R is called closed if R \ is open. Obviously a closed interval [a, b] (where < a b < ) is closed, since the complement is (, a) (b, ), a union of two open sets. 6

Since every (non-empty) open interval contains a rational number and an irrational number, neither Q nor R \ Q is open. 2.10 Lemma Let x R. For each t in some non-empty set T let W t be an open interval containing x. Then V = t T W t is an open interval containing x. Proof. Let A be the inf of V and B the sup. We claim first that A and B are not in V. If B is in V then B is in some W t. Since W t is an open subset of R there is a b > B in W t, and b is in V, which is a contradiction. Thus A and B are not in V and clearly V is a subset of (A, B). We claim that (A, B) is a subset of V. Let y (A, B). Obviously if y = x then y is in V. Suppose that x < y < B. Then (since B is the sup of V ) there is some w with x < y < w such that w lies in some W s. Since x and w lie in the open interval W s, so does y, and y is in V. The same proof works if A < y < x. 2.11 Theorem Let V be a non-empty open subset of R. Then V is the union of countably many pairwise disjoint open intervals. Proof: let x V, and let C x be the union of all open intervals W such that x W V. Then C x is well-defined (since V is open there is at least one such W ) and open, and by Lemma 2.10 this C x is an open interval. We claim that if y is in C x then C x = C y. To see this, note that C x and C y are both open intervals containing y and contained in V, and so is C x C y, by Lemma 2.10. Since x C x C y we get C x C y C x, as C x is the union of all open intervals containing x and contained in V. Since y C x C y, the same argument gives C x C y C y. It follows that if C x and C y have non-empty intersection, with t belonging to both, then both equal C t and so C x = C y. Now let d n be a sequence using up all the rational numbers in V, and put D n = C dn. The set of D n is countable. Since each C x contains a rational number, every C x, for x in V, is one of the D n. So V is the union of the D n. Now we go along our sequence D n, and delete any D n which is equal to one which has occurred previously. 2.12 Lemma (Special case of the Heine-Borel theorem) Let I = [a, b] be a closed interval, and suppose that we have open subsets V 1, V 2, V 3,... of R such that I is contained in the union of the V j. Then I is contained in the union of finitely many 7

of the V j. We remark that this is related to the concept of compactness (G13MTS), but we do not need this concept here. To prove the lemma, suppose the conclusion is false. Then we can find x 1 I \ V 1, x 2 I \ (V 1 V 2 ), and, in general, x n I \ (V 1... V n ). Since a x n b, the sequence (x n ) is bounded, and so has a convergent subsequence x nk, with limit α [a, b]. But then α is is some V N, and there is some u > 0 with (α u, α + u) V N, since V N is open. Since x nk α, we see that for large k we have x nk (α u, α + u) V N. But n k so that for large k we have n k > N and x nk V N V 1... V nk. This contradiction proves the lemma. Again, this is a property of closed intervals [a, b] not shared by other intervals. For example (0, 1] is contained in the union of the open intervals (1/n, 2), n N, and [0, ) is contained in the union of the open intervals ( 1, n), n N. In neither case will finitely many of those intervals suffice to cover the set. 3 Continuous functions 3.1 Basic facts Let be a subset of R and let f : R be a function. We say f is continuous on if the following is true. To each x 0 in and each real ε > 0 corresponds a real δ > 0 such that f(x) f(x 0 ) < ε for all x in with x x 0 < δ. Fact 1: if t n is a sequence in converging to x 0 then f(t n ) converges to f(x 0 ). To see this, if ε > 0 take δ > 0 as above. We have some n 0 such that t n x 0 < δ for all n n 0, giving f(t n ) f(x 0 ) < ε for all n n 0. Fact 2: if is a closed interval [a, b] and f : R is continuous then f has a maximum and minimum on [a, b] and in particular is bounded. To see this, let M be the supremum of the set f() = {f(x) : x }. Take a strictly increasing sequence (y n ) (thus y n < y n+1 ) with limit M. No y n is an upper bound for f(), so we can find s n in f() with y n < s n M. Thus s n tends to M. So there exist t n in such that f(t n ) M. By the Bolzano-Weierstrass theorem we can assume WLOG that the sequence (t n ), being bounded, converges, to x 0 say, and x 0 is in the interval [a, b], since a t n b. Thus f(x 0 ) = M. Hence M is in f() (and M is the max of f()). In particular M is finite. 3.2 Pointwise convergence Let be a subset of R and let f n, n N and f be functions from to R. We say that f n converges pointwise to f on if for each x in, lim f n(x) = f(x). n 8

Thus for each ε > 0 and for each x in, there is an integer N(x) such that f n (x) f(x) < ε for all n N(x). If the f n are continuous, does it follow that f is continuous? The answer is no. xample Let g n (x) be defined on [0, 1] for n N by g n (x) = 1 nx for 0 x 1/n and g n (x) = 0 for x > 1/n. Then g n is continuous on [0, 1]. Set g(x) to be 1 for x = 0 and 0 otherwise. Then g is not continuous, but g n g pointwise on [0, 1]. Notice here also that lim ( lim g n(x)) = 1 lim ( lim g n(x)) = 0. n + x 0+ x 0+ n + A second example displaying the same phenomenon comes from h n (x) = e nx on [0, 1]. Then again h n g pointwise on [0, 1]. So we need a stronger condition which will force the limit function to be continuous. idea is to make N(x) independent of x. The 3.3 Uniform convergence If f n and f are functions from to R we say that f n converges uniformly to f on if the following is true. To each real ε > 0 corresponds an integer N such that f n (x) f(x) < ε for all n N and for all x. The example g n above does not converge uniformly to g on [0, 1]. To see this, take ε = 1/4 and x = 1/2n. Then g(x) = 0 but g n (x) = 1/2. No matter how large we take N, we can choose n N with g n (1/2n) g(1/2n) > ε. 3.4 Theorem If the real-valued functions f n are continuous on and converge uniformly to f : R then f is continuous on. Proof: take x 0 in and ε > 0. Take N so large that f N (x) f(x) < ε/3 for all x in. Take δ > 0 so that f N (x) f N (x 0 ) < ε/3 for all x in with x x 0 < δ. For such x we get f(x) f(x 0 ) f(x) f N (x) + f N (x) f N (x 0 ) + f N (x 0 ) f(x 0 ) < 3ε/3. 4 Riemann Integration The Riemann integral will be defined for continuous and some other functions. It is relatively easy to define and use, and displays the interplay between integration and differentiation well, but it has certain disadvantages. 9

4.1 Basic definitions for the Riemann integral Let f be a bounded real-valued function on the closed interval [a, b] = I. Henceforth a < b unless otherwise explicitly stated. Assume that f(x) M for all x in I. A PARTITION P of I is a finite set x 0,..., x n such that a = x 0 < x 1 <... < x n = b. The points x j are called the vertices of P. We say that partition Q of I is a refinement of partition P of I if every vertex of P is a vertex of Q (i.e. P is a subset of Q). For P as above, we define M k (f) = sup{f(x) : x k 1 x x k } M, m k (f) = inf{f(x) : x k 1 x x k } M. Next, we define the UPPR SUM U(P, f) and LOWR SUM L(P, f) by U(P, f) = n M k (f)(x k x k 1 ), L(P, f) = k=1 n m k (f)(x k x k 1 ). k=1 Notice that L(P, f) U(P, f). The reason we require f to be bounded is so that all the m k and M k are finite and the sums exist. Notice also that M m k M k M for each k, and so M(b a) L(P, f) U(P, f) M(b a). Suppose that f is positive on I and that the area A under the curve exists. It is not hard to see that L(P, f) A U(P, f) for every partition P of I. Further, if you draw for yourself a simple curve, it is not hard to convince yourself that refining P tends to increase L(P, f) and decrease U(P, f). We prove this last statement as a lemma. 4.2 Lemma Let f be a bounded real-valued function on I = [a, b]. (i) If P, Q are partitions of I and Q is a refinement of P, then L(P, f) L(Q, f), U(P, f) U(Q, f). (ii) If P 1 and P 2 are any partitions of I, then L(P 1, f) U(P 2, f). Thus any lower sum is any upper sum. Proof: (i) We first prove this for the case where Q is P plus one extra point. The general case then follows by adding points one at a time. So suppose that Q is the same as P, except that it has one extra vertex c, where x k 1 < c < x k. Then U(Q, f) U(P, f) is (sup{f(x) : x k 1 x c})(c x k 1 ) + (sup{f(x) : c x x k })(x k c) (sup{f(x) : x k 1 x x k })(x k x k 1 ). 10

This is using the fact that all other terms cancel. But the first two sups above are less than or equal to the third. Since c x k 1, x k c, x k x k 1 are all positive, we get U(Q, f) U(P, f) 0. The proof for the lower sums uses the same idea, or can be proved by noting that L(P, f) = U(P, f). (ii) Here we just set P to be the partition obtained by taking all the vertices of P 1 and all those of P 2. We arrange these vertices in order, and P is a refinement of P 1 and of P 2. Now we can write L(P 1, f) L(P, f) U(P, f) U(P 2, f). 4.3 Definition of the Riemann integral Let f be bounded, real-valued on I = [a, b] as before, with f(x) M there. We define the UPPR INTGRAL of f from a to b as b a f(x)dx = inf P {U(P, f)} where the supremum is taken over all partitions P of I. This exists and is finite, because all the upper sums are bounded below by M(b a). Similarly we define the LOWR INTGRAL b a f(x)dx = sup{l(p, f)} P taking the sup over all partitions P of I. Again this exists and is finite, because all the lower sums are bounded above by M(b a). Now we define f to be Riemann integrable on I if the upper integral equals the lower integral, in which case we denote the common value by b a f(x)dx. Notice that the lower integral is always the upper integral, because of Lemma 4.2, part (ii). Also, if f is Riemann integrable and positive on I and the area A under the curve exists, then the fact that L(P, f) A U(P, f) for every partition P of I implies that the lower integral is A and the upper integral is A, which means that A equals b f(x)dx. As usual in integration, it a does not matter whether you write f(x)dx or f(t)dt etc. 4.4 xample Define f on I = [0, 1] by f(x) = 1 if x is rational and f(x) = 0 otherwise. Let P = {x 0,..., x n } be any partition of I. Then clearly M k (f) = 1 for each k, since in each sub-interval [x k 1, x k ] there is a rational number. Thus U(P, f) = n k=1 (x k x k 1 ) = 1 and so the upper integral is 1. Similarly, we have m k (f) = 0 for each k, all lower sums are 0, and the lower integral is 0. Before proving that continuous functions are Riemann integrable, we first deal with the rather easier case of monotone functions (non-decreasing or non-increasing). 11

4.5 Theorem Suppose that f is a monotone function on I = [a, b]. Then f is Riemann integrable on I. Proof: we only deal with the case where f is non-decreasing (i.e. f(x) f(y) for x y). The non-increasing case is similar. Now if f(b) = f(a) then f is constant on I and so the result follows trivially (all upper and lower sums are the same). Assume henceforth that f(b) > f(a). Let ε > 0. We choose a partition P = {x 0,..., x n } such that for each k we have x k x k 1 < ε/(f(b) f(a)). Now, since f is non-decreasing we have M k (f) = f(x k ) and m k (f) = f(x k 1 ). Thus U(P, f) L(P, f) = n (M k (f) m k (f))(x k x k 1 ) = k=1 n (f(x k ) f(x k 1 ))(x k x k 1 ) < k=1 < n (f(x k ) f(x k 1 ))ε/(f(b) f(a)) = (f(x n ) f(x 0 ))ε/(f(b) f(a)) = ε. k=1 Therefore U(P, f) L(P, f) + ε. So the upper integral of f (which is the inf of the upper sums) is at most L(P, f) + ε. But L(P, f) is at most the lower integral (sup of the lower sums). Thus the upper and lower integrals differ by at most ε and, since ε is arbitrary, must be equal. To handle the case of continuous functions, we need the following. 4.6 Uniform continuity Let f be a real-valued function on the closed interval I = [a, b]. We say that f is uniformly continuous on I if the following is true. To each ε > 0 corresponds a δ > 0 such that f(x) f(y) < ε for all x and y in I such that x y < δ. 4.7 Theorem If f is continuous on [a, b] then f is uniformly continuous on [a, b]. Proof: suppose that ε > 0 and that NO positive δ exists with the property in the statement. Then 1/n, for n N, is not such a δ. Thus there are points x n and y n in I with x n y n < 1/n, but with f(x n ) f(y n ) ε. Now (x n ) is a sequence in the closed interval I, and so is a bounded sequence, and therefore we can find a convergent subsequence (x kn ), with limit B, say. Since a x kn b for each n, we have B I. Now x kn y kn 0 as n, and so (y kn ) also converges to B. Since f is continuous on I, we have f(x kn ) f(b) as n and f(y kn ) f(b) as n, which contradicts the fact that f(x kn ) f(y kn ) is always ε. This contradiction proves the theorem. 12

Remark: the name UNIFORM continuity arises because the δ does not depend on the particular choice of x or y. The theorem is NOT true for open intervals, as the example h(x) = 1/x, I = (0, 1) shows. To see this, just note that h(1/n) h(1/(n 1)) = 1 for all n N, but 1/n 1/(n 1) = 1/n(n 1), which we can make as small as we like. 4.8 Theorem Let f be continuous, real-valued, on I = [a, b] (a < b). Then f is Riemann integrable on I. Proof: let ε > 0 be given. We choose a δ > 0 such that for all x and y in I with x y < δ we have f(x) f(y) < ε/(b a). We choose a partition P = {x 0,..., x n } of I such that, for each k, we have x k x k 1 < δ. Now take a sub-interval J = [x k 1, x k ]. We know that there exist c and d in J such that for all x in J we have f(c) f(x) f(d). This means that M k (f) = f(d) and m k (f) = f(c). But c d < δ and so M k (f) m k (f) = f(d) f(c) < ε/(b a). This holds for each k. Thus U(P, f) L(P, f) = n (M k (f) m k (f))(x k x k 1 ) < (ε/(b a)) k=1 The same argument as used for non-decreasing functions now applies. 4.9 Theorem n (x k x k 1 ) = ε. Let f and g be Riemann integrable functions on I. Let c, d be real numbers. Then cf + dg is Riemann integrable on I and b cf(x) + dg(x)dx = c b f(x)dx + d b g(x)dx. a a a Proof: First consider cf, when c > 0. Obviously L(P, cf) = cl(p, f) and U(P, cf) = cu(p, f) and so b cf(x)dx = c b f(x)dx. It is also easy to see that U(P, f) = L(P, f) and a a L(P, f) = U(P, f), so b f(x)dx = b f(x)dx. This leaves only f + g to consider. a a Take ε > 0 and partitions P 1, P 2 of I such that L(P 1, f) and U(P 2, f) are both within ε of b f(x)dx. Note that we also have L(P a 1, f) b f(x)dx U(P a 2, f). We can assume that P 1 = P 2, because otherwise we can replace both by P 1 P 2, and refinements can only push the lower and upper sums closer to the integral. Similarly, take Q such that L(Q, g) and U(Q, g) are both within ε of b g(x)dx. a k=1 We can assume that P = Q, because otherwise we can replace both by P Q. Now L(P, f) + L(P, g) L(P, f + g) U(P, f + g) U(P, f) + U(P, g). Thus L(P, f + g) and U(P, f + g) both lie within 2ε of b f(x)dx + b g(x)dx. Therefore so do a a the upper and lower integrals of f + g and, since ε is arbitrary, the result follows. 13

4.10 The fundamental theorem of the calculus Suppose that F, f are real-valued functions on [a, b], that F is continuous and f is Riemann integrable on [a, b], and that F (x) = f(x) for all x in (a, b). Then b f(x)dx = F (b) F (a). a Proof. The proof is based on the mean value theorem. Let P = {x 0,..., x n } be any partition of [a, b]. Then by the mean value theorem there exist points t k satisfying x k 1 < t k < x k such that F (b) F (a) = n (F (x k ) F (x k 1 )) = k=1 n f(t k )(x k x k 1 ). k=1 But this means that L(P, f) F (b) F (a) U(P, f). Hence the lower integral of f is at most F (b) F (a), and the upper integral of f is at least F (b) F (a). But the lower and upper integrals of f are, by assumption, the same. 4.11 xamples (i) The limit of a sequence of Riemann integrable functions need not be Riemann integrable. Let be the countable set Q [0, 1], and let (r n ) be a sequence in, in which each element of appears exactly once. Define g n as follows. Set g N (x) = 1 if x is one of the N points r 1,..., r N, and g N (x) = 0 otherwise. Obviously all lower sums for g N are 0. Take the partition P = {x 0,..., x n } with x k = k/n, 0 k n, and n > 2N an integer. Then at most 2N intervals [x k 1, x k ] contain a point where g N (x) 0, so U(P, g N ) 2N/n. So 1 g 0 N(x)dx = 0. But as N we see that g N converges pointwise to the function of xample 4.4, which is not Riemann integrable. (ii) Define h n on [0, 1], for integer n > 2, by h n (x) = n 2 x if 0 x 1/n and h n (x) = n 2 (2/n x) if 1/n x 2/n and h n (x) = 0 if x 2/n. Then h n is continuous on [0, 1] with Riemann integral 1. But h n 0 pointwise on [0, 1], and 0 has Riemann integral 0. 5 Series 5.1 The extended real numbers We define R = R {, }. Later we will define some products involving, but for now we just define + =, x + = for all x R. Note that + ( ) is not defined. We can extend < to R R. in the obvious way by saying that < x < for every x in 14

Note that if we say a subset A of R is bounded above this will continue to mean that there is some M R such that x M for all x A: of course all x are also. If A is any non-empty subset of R, then A has a least upper bound (sup) and an inf. Note that sup A is if A has no upper bound in R. In particular this is true if A. We can define limits of sequences in R exactly as in R. In particular x n L R means that given positive real ε we have x n L < ε (and so x n R) for all n n 0 (ε). The monotone sequence theorem remains true: if x n is a non-decreasing sequence in R then x n tends to sup{x n }. 5.2 Series We consider here only series with terms a k in [0, ]. Given such a k for k p N, the partial sums n s n = form a non-decreasing sequence in R, and we set k=p a k S = k=p a k = lim n s n. This will be if any a k is, but of course 1/k =. Note that s n S for each n. k=1 5.3 Re-arrangements Suppose that a k [0, ] for each k N. Suppose that φ : N N is a bijection, and set b k = a φ(k) for each k. Then S 1 = b k = a φ(1) + a φ(2) +......... k=1 is called a R-ARRANGMNT of and the sums S 1, S 2 are equal. S 2 = a k, k=1 15

To see this, just note that if n N, then we can find m such that φ(j) m for 1 j n, and so n n m b k = a φ(k) a p S 2. k=1 k=1 p=1 Letting n we get S 1 S 2. Now just reverse the roles of the two series. Surprisingly, this fails if we allow negative terms. The alternating series S = ln 2 = 1/1 1/2 + 1/3 1/4 +............ re-arranges (one odd, followed by two evens) to 1/1 1/2 1/4 + 1/3 1/6 1/8 + 1/5 1/10 1/12 +....... which has sum S/2. 5.4 Double series Suppose that m and p are (finite) integers and m n and p q and a j,k [0, ] for all integers j, k with m j n, p k q. N.B. j, k do not take the value. Then ( n q ) a j,k = S 1 j=m k=p and are equal. ( q n ) a j,k = S 2 k=p j=m Proof: this is clearly true if n and q are both finite. Also if N n is finite and q =, ( N ) ( ) N M a j,k = a j,k = = lim M j=m j=m ( N M k=p k=p a j,k ) = lim M j=m lim M k=p ( M N ) a j,k = k=p j=m ( N ) a j,k. Setting N = n, this proves the result when n is finite. Now if n and q are both, take any finite N m to get ( N ) ( N ) ( ) a j,k = a j,k a j,k. j=m k=p k=p Letting N we get S 1 S 2, and the reverse inequality is proved the same way. j=m k=p k=p j=m j=m 16

Note that if we take any sum of finitely many a j,k many rows, and so at most the double sum. this will be at most the sum of finitely Again, examples show that this property of independence of order of summation fails if we allow terms of mixed sign. 5.5 Generalized series If T is a countably infinite set, and a t [0, ] for every t in T, we can define t T as follows. Let t n be any sequence using up T, with every t in T appearing exactly once, set b n = a tn and put a t = b n. t T It does not matter which particular sequence we take, because if we use a different sequence the series we get will be a re-arrangement of b n and so have the same sum. a t n=1 For example, 2 t = 2 0 + 2 1 + 2 1 + 2 2 +......... =. t Z 6 Measures We first need: 6.1 σ-algebras Let be a set, and let Π be a collection of subsets of. Then Π is called a σ-algebra if Π is non-empty and the following two conditions are satisfied: (i) for every A in Π, the complement \A is in Π; (ii) if we have countably many A j, say A 1, A 2,..., all in Π, then their union j A j is in Π. In particular, the union of finitely many elements of Π is an element of Π. By taking A ( \ A), we see that is always in Π, and so is the empty set. It follows from (i) and (ii) that the intersection of countably many elements of Π is an element of Π (see problem sheet). The simplest example of a σ-algebra is the power set P (), the collection of all subsets of. Note that some books omit the requirement that Π be non-empty: however, an empty collection of subsets of is not very interesting. Also it is easy to check (optional) that this definition is equivalent to what Dr. Feinstein calls a σ-field in his G1CMIN lecture notes and exam papers. 17

6.2 Lemma Let be a set. If for every t in some set T, the collection Π t is a σ-algebra of subsets of, then the intersection t Π t, which is the collection of all subsets of each belonging to all of the Π t, is itself a σ-algebra of subsets of. The proof is trivial: since Π t for every t, the intersection is non-empty, and conditions (i) and (ii) are obviously satisfied. It follows that every non-empty collection H of subsets of is a sub-collection of a minimal σ-algebra of subsets of. To do this, take the intersection of all σ-algebras of subsets of each of which contain H. This is not a vacuous definition, as there is always at least one, namely P (). This is called the σ-algebra generated by H. The σ-algebra generated by the open sets is called the σ-algebra of Borel sets. We shall see that not every subset of R is a Borel set. 6.3 Measures Let be a non-empty set, and let Π be a (N.B. non-empty) σ-algebra of subsets of. By a measure µ on Π we mean a function µ : Π [0, + ] which satisfies the conditions µ( ) = 0 and µ() = µ( j ) j whenever we have a countable family of pairwise disjoints sets j (all in Π) whose union is. The elements of Π will be called µ-measurable (or just measurable) sets, and we will often talk about µ as a measure on (with the existence of Π taken for granted). 6.4 xamples Let be a set and let µ(u) be the number of elements of each subset U of. Then µ is a measure on P (), called the counting measure. If we have a measure µ on a set and µ() = 1, then µ is called a probability measure. Measurable sets correspond to events. The countable additivity corresponds to the probabilities of pairwise mutually disjoint events. In particular, µ(\u) = 1 µ(u). Let be a set, let x, and let µ(u) be 1 if x U, and 0 otherwise. Then µ is a measure on P () (point mass at x). Let be an uncountable set and, for U, let µ(u) be 0 if U is countable and if U is uncountable. 6.5 Some properties of measures µ (i) If A B then µ(a) µ(b). To see this, write C = B\A = B A c. Then B = A C and A, C are disjoint, and µ(b) = µ(a) + µ(c) µ(a). 18

(ii) µ(a B) = µ(a) + µ(b \ A) µ(a) + µ(b) (using (i)). (iii) µ( n j=1 G j) n j=1 µ(g j). (By induction, using (ii)). (iv) If B j B j+1 and B is the union of the (countably many) B j then µ(b j ) µ(b) as j. To prove (iv), write 1 = B 1 and j+1 = B j+1 \B j. Then B j is the union of 1,..., j (by induction on j) and B is the union of all the j and the j are disjoint. (If m < n we have m B m B n 1 and n = B n \ B n 1 and so m n =. Also if m is the smallest j for which x B j then x m ). We now get µ(b j ) = j µ( m ) m=1 µ( m ) = µ(b). (v) It follows from (iv) that if the F j are all µ-measurable then µ( j=1 F j) = lim n µ( n j=1 F j). (vi) We then have, using (iii) and (v), subadditivity: m=1 µ( j=1 G j ) = lim n µ( n j=1 G j ) lim n n j=1 µ(g j ) = j µ(g j ). (vii) Note that some books omit the condition µ( ) = 0. However µ( ) = µ( ) + µ( ) and so the only possible values for µ( ) are 0 and. If µ( ) = then µ(a) = for every A Π. 7 Lebesgue outer measure We are going to construct the Lebesgue measure λ on a σ-algebra of subsets of R, which may be thought of as a generalization of the idea of the length of an interval. Now the idea of length makes sense for an interval, so that the length of (a, b) is b a (and is if b = or a = ). However, for other sets such as Q the idea of length isn t defined in any obvious way. So we start by constructing the Lebesgue outer measure, which is defined for every subset A of R. The idea is to minimize the total length of open intervals which together cover A. We will then prove some properties, including the fact that for intervals the outer measure equals the length. 7.1 Definition Let A R. It is always possible to choose countably many open intervals which together cover A i.e. whose union contains A: in fact we can do this with one interval (, ). So we define the outer measure λ (A) as follows. Consider all countable collections C of open intervals (a k, b k ) such that A (a k, b k ): these intervals have total length L(C) = k (b k a k ). 19

We do this for all such countable collections C of open sets covering A, and take the inf of L(C) over all possible C. In particular, since A (, + ), this is always defined. Formally, { λ (A) = inf (b k a k ) : A } (a k, b k ), k k in which we consider only covering of A by countably many open intervals. Another way to express this is that λ (A) is the infimum of those positive t such that there exist countably many open intervals (a k, b k ) of total length k (b k a k ) = t with A k (a k, b k ). Obviously, if A B R then any collection of open intervals covering B also covers A, and so λ (A) λ (B). Also {x} (x 1/n, x + 1/n) for n N, so λ ({x}) 2/n and so λ ({x}) = 0. Thus we have λ ( ) = 0. 7.2 Theorem λ is countably sub-additive, which means the following. If we have countably many subsets 1, 2,... of R, and = n n is their union, then λ () n λ ( n ). Proof. This is obvious if any λ ( n ) is infinite. Now assume that all λ ( n ) are finite. Let > δ > 0. For each n, we have λ ( n ) + δ2 n > λ ( n ) and so we can choose a countable family of open intervals I j,n of length L j,n, such that n j I j,n, L j,n < λ ( n ) + δ2 n. j Then is contained in the union of all the I j,n. There are countably many of these intervals I j,n and together they cover. Consider the sum of the lengths of all the I j,n. A partial sum s for this series is the sum of finitely many L j,n. So if N is large enough we get s N L j,n n n=1 j (λ ( n ) + δ2 n ) n λ ( n ) + δ2 n = δ + n n=1 λ ( n ). Since we re taking an arbitrary partial sum, the sum of all the L j,n is at most δ + n λ ( n ). Since δ is arbitrary we get λ () n λ ( n ). 7.3 xamples We have λ (Q) = 0, while λ ([0, 1]\Q) = 1. 20

7.4 Lemma Let A R, let x R, and let A + x = {y + x : y A}. Then λ (A) = λ (A + x). To prove this, let s > λ (A). Then s is not a lower bound for the set of t (0, ] such that A can be covered by the union of countably many open intervals of total length t. So there exists t with t < s such that A is contained in the union of the (a k, b k ), where k (b k a k ) = t. But then A + x is contained in the union of the countably many intervals (a k + x, b k + x), and these have total length t. So λ (A + x) λ (A). The reverse inequality holds, because A = (A + x) + ( x). 7.5 Theorem If A is an interval then λ (A) is the length of A. Proof. Suppose that the interval A has finite end-points a, b. Then for n N we have A (a 1/n, b + 1/n) so λ (A) b a + 2/n. Since n is arbitrary we get λ (A) b a. Thus for a finite interval A, we have λ (A) not more than the length of A, and this is obviously also true for an interval of infinite length. So we need to show that λ (A) is at least the length of A. To prove this, we assume first that A = [a, b], with a, b finite, a < b. Suppose that we have a countable family of open intervals (a k, b k ) (k T N) which together cover A, and k (b k a k ) < b a. By Lemma 2.12 we can assume that there are only finitely many of these intervals. Thus A is contained in the union of N open intervals (a k, b k ). Let n be large and partition A into n closed intervals I 1,..., I n of equal length s = (b a)/n, with vertices in the set {x p = a+ps : p Z}. For each k, the total length of those I j which meet (a k, b k ) is at most b k a k + 2s (it is at most the difference between the least x p which is b k and the greatest x q which is b k ). Since every I j is contained in [a, b] and so meets at least one (a k, b k ), we see that the total length of the I j is (b a) = ns k ((b k a k ) + 2s) 2Ns + k (b k a k ). Since n can be chosen arbitrarily large, with s consequently arbitrarily small, we get b a = ns k (b k a k ). So λ (A) is at least the length of A when A is a closed interval with finite end-points. The general case follows, because if A is an interval and t is positive but less than the length of A, we can choose a closed interval B contained in A, of length t, to get λ (A) λ (B) t. This idea and proof are easily generalized to higher dimensions. In R 2 we take the infimum of the sum (area of B j ), over all coverings of A by open rectangles B j (open means no 21

boundary points included). Consider e.g. the straight line A from (0, 0) to (1, 1). The part with j/n x, y (j + 1)/n lies in the open rectangle (j 1)/n < x, y < (j + 2)/n, which has area 9n 2. So λ (A) 9n 1 and, since n can be arbitrarily large, A has two-dimensional outer measure 0. 8 Lebesgue measure The outer measure λ of the last section turns out not to be a measure on the whole power set P (R). However, we can find a σ-algebra of subsets of R on which it is a measure. The idea is the following. If λ were a measure on the power set of R then we d have λ (A) = λ (A ) + λ (A\) for every A,, by disjointness. So we consider those for which this is true for every A. 8.1 Definition A subset of R is said to be Lebesgue measurable if we have for every subset A of R. Note that λ (A) = λ (A ) + λ (A\) λ (A) λ (A ) + λ (A\) always holds, so that to show that some set is Lebesgue measurable it suffices to prove that for every A. We will sometimes write c for R \. 8.2 Some basic facts λ (A) λ (A ) + λ (A\) (i) R and the empty set are Lebesgue measurable. (ii) Any set with λ () = 0 is Lebesgue measurable. (iii) If is Lebesgue measurable, so is R\. (iv) If is Lebesgue measurable, so is x + for every x R. Property (iii) is obvious. To prove (ii), assume λ () = 0. Then λ (A ) will be 0 for all A, which gives λ (A ) + λ (A \ ) = λ (A \ ) λ (A) 22

as required. Property (i) now follows from (ii) and (iii). Now we prove (iv). Let R be Lebesgue measurable, and note that for A R we have A ( + x) = {y : y A, y x } = {z + x : z + x A, z } = = x + {z : z A x, z } = x + ((A x) ) and hence λ (A ( + x)) = λ ((A x) ), using Lemma 7.4. Next, note that ( + x) c = c + x. This gives A \ ( + x) = A ( + x) c = A ( c + x) = x + ((A x) c ) which gives λ (A \ ( + x)) = λ ((A x) c ). Since is Lebesgue measurable we now get λ (A) = λ (A x) = λ ((A x) ) + λ ((A x) c ) = λ (A ( + x)) + λ (A \ ( + x)). 8.3 Theorem The union or intersection of finitely many Lebesgue measurable sets is Lebesgue measurable. We first prove that if and F are Lebesgue measurable subsets of R, then so is G = F. Write c for R\. We have λ (A) = λ (A )+λ (A c ) = λ (A F )+λ (A F c )+λ (A c F )+λ (A c F c ) = = λ (A F ) + λ (A F c ) + λ (A c F ) + λ (A G c ) λ (A G) + λ (A G c ), since λ is sub-additive and G = F = ( F ) ( F c ) ( c F ). Now, if we are given n 3 and Lebesgue measurable sets F 1,... F n, we write n F j = F n j=1 and so the assertion about unions of finitely many sets follows by induction. Intersections work as well, because n is the complement of j=1 n F j n 1 j=1 F j and each F c j is Lebesgue measurable. j=1 F c j 23

8.4 Theorem (i) Suppose that F 1, F 2,... are countably many pairwise disjoint Lebesgue measurable sets with union F. Then F is Lebesgue measurable, and λ (F ) = j λ (F j ). (ii) The union of countably many Lebesgue measurable sets is Lebesgue measurable. Proof. (i) We may assume that there are infinitely many F j, by making them all empty from some j on, if necessary. Let G n = n j=1 F j. The G n are Lebesgue measurable. We claim that for every n N, and for every A R, λ (A G n ) = n λ (A F j ). j=1 This is clearly true for n = 1. Assuming it true for n and using the Lebesgue measurability of F n+1, we get n+1 λ (A G n+1 ) = λ (A G n+1 F n+1 )+λ (A G n+1 Fn+1) c = λ (A F n+1 )+λ (A G n ) = λ (A F j ) and the result follows. Now, by the previous claim, if A is any subset of R, j=1 λ (A) = λ (A G n ) + λ (A G c n) λ (A G n ) + λ (A F c ) = n λ (A F j ) + λ (A F c ). j=1 Since n is arbitrary, λ (A) j λ (A F j ) + λ (A F c ). (2) Since λ is countably sub-additive, we get, from (2), λ (A) λ (A F ) + λ (A F c ), which proves that F is Lebesgue measurable. Now choosing A = F in (2) and using sub-additivity again gives λ (F ) λ (F j ) λ (F ). j To prove (ii), just note that if we have 1, 2,... with union, then setting H 1 = 1, H n+1 = n+1 \( n j=1 j), the H j are Lebesgue measurable and pairwise disjoint, and their union is. (If m < n then H m m and H n m is empty). 8.5 Corollary For Lebesgue measurable F we define λ(f ) = λ (F ). The Lebesgue measurable sets form a σ-algebra Π of subsets of R, and λ is a measure on Π. 24

8.6 Theorem Let a R. Then the open interval (a, + ) is Lebesgue measurable. Proof. Let A be any subset of R, and let A 1 = A (a, + ), A 2 = A (, a]. We only need to show that λ (A) λ (A 1 ) + λ (A 2 ), which is obvious if λ (A) =. Assume now that λ (A) is finite. Take a real δ > 0 and choose a countable collection of open intervals I j which cover A, such that I j < λ (A) + δ, using I for the length. Let j P j = I j (a, + ), Q j = I j (, a]. Then A 1 j P j so λ (A 1 ) j λ (P j ) = j P j since P j is an interval. Doing the same for A 2, λ (A 1 ) + λ (A 2 ) j P j + j Q j = j I j < λ (A) + δ. Since δ is arbitrary the theorem is proved. 8.7 Corollary All Borel sets (in particular, all open sets) are Lebesgue measurable. 8.8 Theorem Let R. The following statements are equivalent. (i) is Lebesgue measurable. (ii) For every real ε > 0 there exists an open set U with U and λ (U \ ) < ε. (iii) There exists a set V, such that: V is the intersection of countably many open sets; V contains ; λ (V \ ) = 0. Thus Lebesgue measurable sets may in a certain sense be well-approximated by open sets. Proof. (ii) implies (iii): for each n N take an open set V n containing, such that λ (V n \) < 1/n. Now let V be the intersection of the V n. Then V, and for each n we have λ (V \ ) λ (V n \ ) < 1/n. (iii) implies (i): take V as in (iii), and set W = R\V. Then W F = R\, and F \W = V \. 25

Since V is the intersection of open U j, say, then W = R \ U j = (R \ U j ) is a union of closed sets, and so is Lebesgue measurable. Hence we can write F as the union of a closed set W and a set F \ W which has outer measure 0. Thus F is Lebesgue measurable and so is. (i) implies (ii). First set n = [ n, n], n N. ach n is Lebesgue measurable. For each n N choose, directly from the definition of outer measure, a countable union U n of open intervals containing n and having sum of lengths less than λ( n ) + ε2 n. Then A n = U n \ n has λ( n ) + λ(a n ) = λ(u n ) and so λ(a n ) = λ (A n ) < ε2 n. Let U be the union of the U n, n N. Then U is open, and U \ = ( U n ) \ = (U n \ ) (U n \ n ) = A n has, by the subadditivity of λ, outer measure at most n N λ (A n ) < ε(1/2 + 1/4 +...) = ε. 9 A set which is not Lebesgue measurable 9.1 The Axiom of Choice The version of this axiom which we will use is the following: Suppose that we have a set T, and that A t is a non-empty set, for each t T, and that A t A s =, for s, t T, s t. Then we can form a set B = {c t : t T } by choosing one c t from each A t. 9.2 Theorem There exists a subset of [0, 1] such that is not Lebesgue measurable. Proof. We define a relation on [0, 1] by x y iff x y is rational. Then is an equivalence relation. To see this, obviously x x (so is reflexive) and x y iff y x (symmetric) and x y and y z imply that y x and z y are rational, so that z x is rational (transitive). For each x I = [0, 1] we form the equivalence class [x] = {y I : y x}. Then either [x] [y] is empty, or [x] = [y]. Thus I is the union of these disjoint equivalence classes. We have a set of pairwise disjoint equivalence classes, whose union is I. (To see that we have a set T of these, use the mapping φ : I T given by φ(x) = [x], so that T is just the image φ(i).) Using the Axiom of Choice, we form a set which contains precisely one element of each equivalence class. So for each x in [0, 1] there is a unique y in such that x y is rational. Note that 1 x y 1. Now use the fact that H = Q [ 1, 1] is countable, and write this set as {r 1, r 2, r 3,...}, with the r j distinct rational numbers, using up H. Then every x in [0, 1] belongs to one of the j defined by j = + r j. Also, if j k then j k =. For if u lies in both then u r j and u r k are both in. But (u r j ) (u r k ) and (u r j ) (u r k ), and contains just one element of each equivalence class. 26

Suppose that is Lebesgue measurable. Then so is each j. Now, [0, 1] is contained in the countable union of the pairwise disjoint j, so 1 = λ([0, 1]) = λ ([0, 1]) j N λ ( j ) = j N λ( j ) = j N λ(). So λ() > 0. But the j are disjoint and each is a subset of [ 1, 2], so = j N which is impossible. Hence it is not always true that λ() = j N λ( j ) λ([ 1, 2]) = 3, λ (A) = λ (A ) + λ (A c ), so the Lebesgue outer measure described earlier is not additive on the whole power set of R, and so is not a measure on the power set of R. We also see that not every subset of R is a Borel set. Not all mathematicians accept the Axiom of Choice. It can be shown to imply that every set C has an ordering < with the properties that, for all a, b, c in C, (i) either a < b or b < a or b = a, and a < b and b < c implies a < c, as well as (ii) every non-empty subset D of C has a least element i.e. there exists d D such that d < c for all c in D with c d. Such an ordering is called a well-ordering. N is well-ordered by ordinary <, but R is not. very countable set A has an obvious well-ordering. Write A as {a j } without repetition, and order by a 1 < a 2 < a 3 <.... 10 Measurable functions We have constructed Lebesgue measure for (some) subsets of R, and we now return to a general measure space. So assume we have a non-empty set, a (non-empty) σ-algebra Π of subsets of, and a measure µ : Π [0, ]. We will sometimes refer to the elements of Π as µ-measurable sets. The aim is to construct the integral fdµ for measurable, and to develop its properties, and in this chapter we determine which functions can be used. Since we occasionally need products of functions, we define some products involving. 10.1 Products involving infinity We now define if x > 0 and x. = 0. = 0. The purpose of the latter is to make certain integrals take their expected values. 27