Lebesgue-Radon-Nikodym Theorem

Similar documents
5 Measure theory II. (or. lim. Prove the proposition. 5. For fixed F A and φ M define the restriction of φ on F by writing.

HILBERT SPACES AND THE RADON-NIKODYM THEOREM. where the bar in the first equation denotes complex conjugation. In either case, for any x V define

Signed Measures. Chapter Basic Properties of Signed Measures. 4.2 Jordan and Hahn Decompositions

Signed Measures and Complex Measures

Real Analysis, 2nd Edition, G.B.Folland Signed Measures and Differentiation

FUNDAMENTALS OF REAL ANALYSIS by. IV.1. Differentiation of Monotonic Functions

Measure Theory on Topological Spaces. Course: Prof. Tony Dorlas 2010 Typset: Cathal Ormond

Probability and Random Processes

MATHS 730 FC Lecture Notes March 5, Introduction

Chapter 8. General Countably Additive Set Functions. 8.1 Hahn Decomposition Theorem

2 Lebesgue integration

( f ^ M _ M 0 )dµ (5.1)

212a1214Daniell s integration theory.

1 Inner Product Space

Lecture 5 Theorems of Fubini-Tonelli and Radon-Nikodym

36-752: Lecture 1. We will use measures to say how large sets are. First, we have to decide which sets we will measure.

Problem Set. Problem Set #1. Math 5322, Fall March 4, 2002 ANSWERS

INTRODUCTION TO MEASURE THEORY AND LEBESGUE INTEGRATION

LEBESGUE MEASURE AND L2 SPACE. Contents 1. Measure Spaces 1 2. Lebesgue Integration 2 3. L 2 Space 4 Acknowledgments 9 References 9

MATH MEASURE THEORY AND FOURIER ANALYSIS. Contents

CHAPTER 6. Differentiation

Defining the Integral

Section Signed Measures: The Hahn and Jordan Decompositions

Section The Radon-Nikodym Theorem

Proof of Radon-Nikodym theorem

Exercise 1. Show that the Radon-Nikodym theorem for a finite measure implies the theorem for a σ-finite measure.

l(y j ) = 0 for all y j (1)

4 Integration 4.1 Integration of non-negative simple functions

THEOREMS, ETC., FOR MATH 515

The Caratheodory Construction of Measures

Analysis Comprehensive Exam Questions Fall 2008

Real Analysis Chapter 3 Solutions Jonathan Conder. ν(f n ) = lim

Annalee Gomm Math 714: Assignment #2

(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.

MATH 650. THE RADON-NIKODYM THEOREM

1 Measurable Functions

Math 4121 Spring 2012 Weaver. Measure Theory. 1. σ-algebras

02. Measure and integral. 1. Borel-measurable functions and pointwise limits

Differentiation of Measures and Functions

Real Analysis Notes. Thomas Goller

+ 2x sin x. f(b i ) f(a i ) < ɛ. i=1. i=1

Chapter 4. Measure Theory. 1. Measure Spaces

AN INTRODUCTION TO GEOMETRIC MEASURE THEORY AND AN APPLICATION TO MINIMAL SURFACES ( DRAFT DOCUMENT) Academic Year 2016/17 Francesco Serra Cassano

MAT 571 REAL ANALYSIS II LECTURE NOTES. Contents. 2. Product measures Iterated integrals Complete products Differentiation 17

Three hours THE UNIVERSITY OF MANCHESTER. 24th January

MTH 404: Measure and Integration

Stanford Mathematics Department Math 205A Lecture Supplement #4 Borel Regular & Radon Measures

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

Integration on Measure Spaces

If Y and Y 0 satisfy (1-2), then Y = Y 0 a.s.

QUANTUM MEASURE THEORY. Stanley Gudder. Department of Mathematics. University of Denver. Denver Colorado

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

Chapter 6. Integration. 1. Integrals of Nonnegative Functions. a j µ(e j ) (ca j )µ(e j ) = c X. and ψ =

3 Integration and Expectation

18.175: Lecture 3 Integration

2 Measure Theory. 2.1 Measures

M17 MAT25-21 HOMEWORK 6

REAL ANALYSIS I Spring 2016 Product Measures

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

Real Analysis Problems

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

L p Functions. Given a measure space (X, µ) and a real number p [1, ), recall that the L p -norm of a measurable function f : X R is defined by

THEOREMS, ETC., FOR MATH 516

Lebesgue Measure on R n

Lecture 1 Real and Complex Numbers

CHAPTER I THE RIESZ REPRESENTATION THEOREM

Dynkin (λ-) and π-systems; monotone classes of sets, and of functions with some examples of application (mainly of a probabilistic flavor)

Measure and integration

Measure Theory. John K. Hunter. Department of Mathematics, University of California at Davis

Folland: Real Analysis, Chapter 7 Sébastien Picard

MATH & MATH FUNCTIONS OF A REAL VARIABLE EXERCISES FALL 2015 & SPRING Scientia Imperii Decus et Tutamen 1

Measure Theory & Integration

Homework 11. Solutions

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

Lebesgue measure and integration

Measurable functions are approximately nice, even if look terrible.

Notes on Measure, Probability and Stochastic Processes. João Lopes Dias

MA359 Measure Theory

ABSTRACT INTEGRATION CHAPTER ONE

A List of Problems in Real Analysis

MATH 418: Lectures on Conditional Expectation

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

Reminder Notes for the Course on Measures on Topological Spaces

Construction of a general measure structure

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

Solutions to Tutorial 11 (Week 12)

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

Analysis of Probabilistic Systems

Review of measure theory

Chapter IV Integration Theory

Homework 4, 5, 6 Solutions. > 0, and so a n 0 = n + 1 n = ( n+1 n)( n+1+ n) 1 if n is odd 1/n if n is even diverges.

Riesz Representation Theorems

9 Radon-Nikodym theorem and conditioning

Section Integration of Nonnegative Measurable Functions

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

Class Notes for Math 921/922: Real Analysis, Instructor Mikil Foss

Measures and Measure Spaces

LECTURE NOTES FOR , FALL Contents. Introduction. 1. Continuous functions

Notions such as convergent sequence and Cauchy sequence make sense for any metric space. Convergent Sequences are Cauchy

Final. due May 8, 2012

Transcription:

Lebesgue-Radon-Nikodym Theorem Matt Rosenzweig 1 Lebesgue-Radon-Nikodym Theorem In what follows, (, A) will denote a measurable space. We begin with a review of signed measures. 1.1 Signed Measures Definition 1. A signed measure ν on a σ-algebra A is a set function such that 1. ν is extended-valued in the sense that < ν() for all A. If { j } are disjoint subsets of A, then ν j = ν( j ) Note that in order for condition () to hold, the sum ν( j ) must be independent of any rearrangement. If ν( j) is finite, then it follows from a theorem of Riemann that the sum converges absolutely. I have included a proof of this result for the interested reader. Lemma. (Riemann Rearrangement Lemma) Let a n be a series of real numbers which converges, but not absolutely. Suppose α β. Then there exists a rearrangement a n with partial sums s n such that Proof. For n N, set lim inf s n = α, lim sup s n = β p n = a n + a n, q n = a n a n Then p n q n = a n, p n + q n = a n, p n 0, q n 0. I claim that the series p n, q n diverge. If either pn or p n converges, then since a n converges, it follows that a n converges, which contradicts our hypothesis that a n does not converge absolutely. Let P 1, P, denote the nonnegative terms of a n, in the order in which they occur, and let Q 1, Q, denote the absolute values of the negative terms of a n also in their original order. Since N n=1 P n = MN n=1 p n and N n=1 Q n = M N n=1 q n for all N N, both P n, Q n diverge. We will construct sequences (m n ) n=1, (k n ) n=1, such that P 1 + + P m1 Q 1 Q k1 + P m1+1 + + P m Q k1+1 Q k + satisfies the conclusion of the lemma. Choose real sequences (α n ) n=1, (β n ) n=1 such that α n α, β n β. Let m 1, k 1 be the minimal positive integers such that and Let m, k be the minimal positive integers such that P 1 + + P m1 > β 1, P 1 + + P m1 Q 1 Q k1 < α 1 P 1 + + P m1 Q 1 Q k1 + P m1+1 + + P m > β, and P 1 + + P m1 Q 1 Q k1 + P m1+1 + + P m Q k1+1 Q k < α 1

We continue in this fashion. This selection process is possible since P n, Q n diverge. Let x n, y n denote the partial sums whose last terms are P mn and Q kn, respectively. Then x n b n P mn, y n α n Q kn Since a n is convergent, P mn, Q kn 0 as n. We conclude that x n β, y n α. Since the subsequential limits of the rearrangement series a n are bounded from above by β and bounded from below by α (this is evident from the squeeze theorem), it follows immediately that lim sup s n = β, lim inf s n = α Corollary 3. If every rearrangement of real-valued series n=1 a n converges, then n=1 a n <. 1. Total Variation Given a signed measure ν on a measure space (, A), one might ask if it is always possible to find a positive measure µ which dominates ν in the following sense: ν() µ(), A Moreover, one might ask if there is a smallest such µ in the sense that if µ is any other positive measure which dominates ν, then µ() µ () for all A. It turns out the answer is yes. First, we need to define the notion of the total variation of a measure, analogous to the variation of a measurable function. Definition 4. Define a set function ν : A R, called the total variation of ν, by ν () = sup ν( j ) where the supremum is taken over all countable partitions of. Lemma 5. The total variation ν of a signed measure ν is a positive measure which satisfies ν nu. Proof. The only axiom of measures which ν does not obviously satisfy is σ-additivity. Let { j } be a countable collection of disjoint sets in A, and set = j. For each j, let α j R such that α j < v ( j ). It follows from the definition of supremum and of v that, for each j, there exists a countable collection of disjoint sets {F i,j } i=1 such that j = i=1 F i,j and Since {F i,j } i, is a partition of, we have α j ν(f i,j ) i=1 α j ν(f i,j ) ν () i=1 Taking the supremum over all α j which satisfy α j < v ( j ) yields the inequality ν ( j ) ν () For the reverse inequality, let {F k } be any other partition of. For k fixed, {F k j } is a partition of F k. By the σ-additivity of ν, ν(f k ) = ν(f k j )

If ν(f k j ) =, then ν(f k j ) =. It follows from the definition of total variation that ν ( j) =, and the inequality ν(f k ) ν () holds trivially. If ν(f k j ) <, then we can interchange the order of summation to obtaion ν(f k ) = ν(f k j ) = ν(f k j ) ν ( j ), since {F k j } is a partition of j for each fixed j. Since {F k } obtain the reverse inequality ν () ν ( j ), which completes the proof. was an arbitrary partition of, we Analogous to the decomposition of a function f into a difference f = f + f of two nonnegative functions, we can write a signed measure as the difference of two positive measures. Definition 6. For a signed measure ν, we define the positive variation and negative variation of ν by ν + = 1 ( ν + ν) and ν = 1 ( ν ν) If A is such that ν() =, then ν () := 0. It follows from the preceding proposition that ν +, ν are both (positive) measures, and moreover, ν = ν + ν and ν = ν + + ν We say that the signed measure ν is σ-finite if the measure ν is σ-finite. 1.3 Mutual Singularity and Absolute Continuity Definition 7. Two signed measures ν and µ on a measure space (, A) are mutually singular if there are disjoint subsets A, B A so that ν() = ν(a ) and µ() = µ(b ), A If ν, µ are mutually singular, we write ν µ. If ν is a signed measure and ν is a positive measure on A, we say that ν is absolutely continuous with respect to µ if A, µ() = 0 ν() = 0 If ν is absolutely continuous with respect to µ, we write ν µ. Lemma 8. If ν and µ are mutually singular, and ν is also absolutely continuous with respect to µ, then ν vanishes identically. Proof. Let A, B A be disjoint subsets such that ν() = ν(a ) and µ() = µ(b ), A For any A, µ(a ) = 0. By absolute continuity, ν(a ) = ν() = 0. Since A was arbitrary, we conclude that ν = 0. Lemma 9. Let µ be a positive measure and ν be a signed measure. 1. If for every ɛ > 0, there exists δ > 0 such that A, µ() < δ ν() < ɛ, then ν µ. 3

. If ν is a finite measure, then the converse holds. Proof. We first prove (1). Let A be such that µ() = 0. It follows from our hypothesis that ν() < ɛ for every ɛ > 0, from which we conclude that ν() = 0. We prove () by contradiction. Suppose there exists ɛ > 0 such that for all δ > 0, there exists δ A with µ( δ ) < δ and ν( δ ) ɛ. For each n N choose n A with µ( n ) 1 and ν( n n ) ɛ. Then ( µ lim sup ) n = µ n=1 k n for all N N. We conclude that µ(lim sup n ) = 0. Since ν µ by hypothesis, we have ν(lim sup n ) = 0. But this is a contradiction since ν(lim sup k k=n n ) = lim sup ν( n ) ɛ 1 k 1.4 Lebesgue-Radon-Nikodym Theorem There are several proofs of the (Lebesgue-)Radon-Nikodym theorem, but I am fond of the one given below because it uses the theory of Hilbert spaces, which is quite elegant. The exposition closely follows that of Stein and Shakarchi in Real Analysis: Measure, Integration, and Hilbert Spaces. Theorem 10. Suppose µ is a σ-finite positive measure on the measure space (, A) and ν is a σ-finite signed measure on A. Then there exist unique signed meaures ν a and ν s on mathcala such that ν a µ, ν s µ, and ν = ν s + ν a. Furthermore, the measure ν a is given by ν a () = f(x)µ(dx), A for some extended µ-integrable function f. Proof. We first consider the case where both µ and ν are positive and finite. Set ρ = µ + nu. We define a functional l : L (, ρ) C by l(ψ) = ψ(x)ν(dx) Clearly, l is linear. I claim that l is bounded. Indeed, since ν, µ are both positive, ( l(ψ) ψ(x) ν(dx) ψ(x) ρ(dx) = ψ(x)1 (x) ρ(dx) (ρ()) 1 ) 1 ψ(x) ρ(dx), where the last inequality follow from the Cauchy-Schwarz inequality. Since L (, ρ) is a Hilbert space, the Riesz representation theorem tells us that there exists a unique (up to a.e. equivalence) g L (, ρ) such that ψ(x)ν(dx) = ψ(x)g(x)ρ(dx), ψ L (, ρ) If A with ρ() > 0, when we set ψ = χ and recall that ν ρ, we obtain 0 1 g(x)ρ(dx) = ν() ρ() ρ() ρ() ρ() = 1, I claim that 0 g(x) 1 ρ-a.e. Indeed, 0 g(x)ρ(dx) for all sets A implies that 0 g(x)ρ(dx) 1 { n ρ g < 1 } { ρ g < 1 } = 0, n N n n g< 1 n Taking the intersection yields ρ {g < 0} = 0. By the same argument, 0 (1 g(x))ρ(dx) for all A implies that g(x) 1 ρ-a.e. Thus, we may assume that 0 g(x) 1 for all x, and we have ψ(1 g)dν = ψgdµ 4

Consider the two sets and define two measures ν a and ν s on A by A = {x : 0 g(x) < 1} and B = {x : g(x) = 1} ν a () = ν(a ) and ν s () = ν(b ), A The diligent inclined reader can verify that ν a, ν s are indeed measures. I claim that ν s µ. It is tautological that A, B are disjoint and ν s () = 0 for all measurable subsets A, so we need only show that µ() = 0 for all measurable subsets B. Indeed, taking ψ = 1 in the identity ψgdµ = ψ(1 g)dν yields µ() = 1 gdµ = 1 (1 g)dµ = 1 0dµ = 0 I now claim that ν a µ. Let A be such that µ() = 0. Then 0 = gdµ = (1 g)dν Since 1 g 0, we conclude that (1 g)1 = 0 a.e., which imples that ν a () = ν( A) = 0. I now claim that dv a = fdµ. Let A, and set ψ = ( n k=0 gk )1. Then (1 g n+1 )dν = n+1 gdµ If x B, then (1 g n+1 )(x) = 0, and if x A, then (1 g n+1 )(x) 1, n. In other words, lim 1 g n+1 = 1 A. Since 1 g n+1 and our measure space is finite, the dominated convergence theorem implies that lim (1 g n+1 )dν = 1 A dν = ν(a ) = ν A () Observe that { n+1 g(x) lim g k 1 g(x) x A (x) = x B Set f = g 1 g. From the monotone convergence theorem we conclude that n+1 v a () = lim g k dµ = Furthermore, f L 1 (, µ) since fdµ = ν a() ν() <. We now consider the case where ν, µ are σ-finite positive measures. It follows from the definition of σ-finite that we can find pairwise disjoint sets j A such that = j and µ( j ), ν( j ) < for all j. We define positive, finite measures on A by fdµ µ j () = µ( j ) and ν j () = ν( j ), A For each j, we write ν j = ν j,a + ν j,s, where ν j,s µ j and ν j,a = f j dµ j. Defining f = j f j, ν s = j ν j,s, ν a = j ν j,a completes the argument. If ν is signed, then we apply the preceding argument separately to the positive and negative variations of ν. To see the uniqueness of the decomposition, suppose we also have ν = ν a + ν s, where ν a µ and ν s µ. Then ν a ν a = ν s ν s 5

Clearly, the LHS is absolutely continuous with respect to µ. I claim that the RHS is singular with respect to µ. Indeed, let A B, A B be paritions guaranteed in the definition of singular measures. Then (ν s ν s ) () = 0 for all measurable subset A A, and for any \ (A A ), µ() = µ( B \ B ) + µ( B \ B) + µ( B B ) = 0 + 0 + 0 = 0 We conclude that ν a ν a = 0 = ν s ν s. 6