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

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

Signed Measures and Complex Measures

Examples of Dual Spaces from Measure Theory

CHAPTER 6. Differentiation

MATHS 730 FC Lecture Notes March 5, Introduction

6.2 Fubini s Theorem. (µ ν)(c) = f C (x) dµ(x). (6.2) Proof. Note that (X Y, A B, µ ν) must be σ-finite as well, so that.

Recall that if X is a compact metric space, C(X), the space of continuous (real-valued) functions on X, is a Banach space with the norm

Section Signed Measures: The Hahn and Jordan Decompositions

Integration on Measure Spaces

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

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

Chapter 4. Measure Theory. 1. Measure Spaces

1.4 Outer measures 10 CHAPTER 1. MEASURE

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

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.

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

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

The Caratheodory Construction of Measures

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

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

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

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

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

Differentiation of Measures and Functions

MATH 650. THE RADON-NIKODYM THEOREM

Lebesgue-Radon-Nikodym Theorem

3. (a) What is a simple function? What is an integrable function? How is f dµ defined? Define it first

2 Measure Theory. 2.1 Measures

(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.

Chapter 4. Measurable Functions. 4.1 Measurable Functions

for all x,y [a,b]. The Lipschitz constant of f is the infimum of constants C with this property.

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

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

i. v = 0 if and only if v 0. iii. v + w v + w. (This is the Triangle Inequality.)

Riesz Representation Theorems

INTRODUCTION TO MEASURE THEORY AND LEBESGUE INTEGRATION

CHAPTER I THE RIESZ REPRESENTATION THEOREM

+ = x. ± if x > 0. if x < 0, x

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

Analysis Comprehensive Exam Questions Fall 2008

Homework 11. Solutions

Measure and integration

1.1. MEASURES AND INTEGRALS

ABSTRACT INTEGRATION CHAPTER ONE

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

A List of Problems in Real Analysis

Disintegration into conditional measures: Rokhlin s theorem

2 Lebesgue integration

MATH MEASURE THEORY AND FOURIER ANALYSIS. Contents

MTH 404: Measure and Integration

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

The Lebesgue Integral

The Hahn-Banach and Radon-Nikodym Theorems

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

THEOREMS, ETC., FOR MATH 516

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 11 (Week 12)

2 (Bonus). 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?

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

Lebesgue measure and integration

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

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?

( f ^ M _ M 0 )dµ (5.1)

L p Spaces and Convexity

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

Real Analysis Problems

THEOREMS, ETC., FOR MATH 515

Measure Theory, 2009

Real Analysis Chapter 1 Solutions Jonathan Conder

Measures and Measure Spaces

Partial Solutions to Folland s Real Analysis: Part I

Probability and Measure

Folland: Real Analysis, Chapter 7 Sébastien Picard

REAL ANALYSIS LECTURE NOTES: 1.4 OUTER MEASURE

Analysis of Probabilistic Systems

Defining the Integral

Real Analysis, 2nd Edition, G.B.Folland Elements of Functional Analysis

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

Basics of Stochastic Analysis

MATH 418: Lectures on Conditional Expectation

Three hours THE UNIVERSITY OF MANCHESTER. 24th January

Measure Theory & Integration

Construction of a general measure structure

PRODUCT MEASURES AND FUBINI S THEOREM

Chapter IV Integration Theory

MA359 Measure Theory

Dual Space of L 1. C = {E P(I) : one of E or I \ E is countable}.

REAL ANALYSIS I Spring 2016 Product Measures

I. ANALYSIS; PROBABILITY

Real Analysis Notes. Thomas Goller

MATH 202B - Problem Set 5

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

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

Chapter 4. The dominated convergence theorem and applications

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

Chapter 5. Measurable Functions

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

Lecture 10. Theorem 1.1 [Ergodicity and extremality] A probability measure µ on (Ω, F) is ergodic for T if and only if it is an extremal point in M.

Probability and Random Processes

Math 5051 Measure Theory and Functional Analysis I Homework Assignment 2

Transcription:

Chapter 8 General Countably dditive Set Functions In Theorem 5.2.2 the reader saw that if f : X R is integrable on the measure space (X,, µ) then we can define a countably additive set function ν on by the formula ν() = f dµ (8.1) and we see that it can take both positive and negative values. In this chapter we will study general countably additive set functions that can take both positive and negative values. Such set functions are known also as signed measures. In the Radon-Nikodym theorem we will characterize all those countably additive set functions that arise from integrals as in Equation 8.1 as being absolutely continuous with respect to the measure µ. nd in the Lebesgue Decomposition theorem we will learn how to decompose any signed measure into its absolutely continuous and singular parts. These concepts for signed measures will be defined as part of the work of this chapter. 8.1 Hahn Decomposition Theorem Definition 8.1.1. Given a σ-algebra of subsets of X, a function µ : R is called a countably additive set function (or a signed measure 1 ) provided 1 Some authors allow a signed measure to be extended real-valued. In that case it is necessary to require that µ take only one of the two values ± in order to ensure that µ is well-defined on. 131

that for every sequence of mutually disjoint sets n we have ( ) µ = µ( n ). n N n N n We prove first the following theorem. Theorem 8.1.1. If µ is a countably additive set function on a σ-field, then µ is bounded on. That is, there exists a real number M such that µ() M for all. Proof. We begin by restating the theorem as follows, bearing in mind that µ can have both positive and negative values on, and that consequently µ need not be monotone. Let µ () = sup{ µ(b) B, B } for each. Then the theorem claims that µ (X) <. i. We claim that both µ() and µ () are sub-additive as functions of 2. The first inequality follows immediately from the triangle inequality for real numbers, combined with the additivity of µ: If and B are -measurable and disjoint, then µ ( B) = µ() + µ(b) µ() + µ(b). For the second inequality, we note that if = 1 2, a disjoint union, then µ () µ ( 1 ) + µ ( 2 ) because if an -measurable set B then µ(b) µ(b 1 ) + µ(b 2 ) µ ( 1 ) + µ ( 2 ) where we have used in the first inequality the sub-additivity of µ. ii. We will suppose that µ (X) = and deduce a contradiction. By hypothesis, µ(x) R. So there exists a set B such that µ(b) > µ(x) + 1 1. 2 We do not denote µ() in the form µ () because the latter symbol will be given a special meaning in Definition 8.1.2. 132

By the additivity of µ, µ(x \ B) = µ(x) µ(b) µ(b) µ(x) > 1. Because B and X \ B are disjoint, it follows from sub-additivity that either µ (B) = or µ (X \ B) =. Thus there exists B 1 such that µ(b 1 ) > 1 and µ (X \ B 1 ) =. Hence there exists B 2, disjoint from B 1, such that µ(b 2 ) > 1 and µ (X \ (B 1 B 2 )) =. This process generates an infinite sequence of mutually disjoint sets B n such that µ(b n ) > 1 for each n N. Let B = n N B n so that µ(b) = n N µ(b n ) (8.2) and the latter series is conditionally convergent, meaning that it is convergent, but not absolutely convergent. Therefore, by a familiar exercise or theorem from dvanced Calculus 3, both the sum of the positive terms and the sum of the negative terms in Equation 8.2 must diverge. Hence there exists a subsequence B nj such that ( ) µ B nj / R which is a contradiction. j N We are ready now to state and prove the Hahn Decomposition theorem. Theorem 8.1.2. (Hahn Decomposition) Let µ be a countably additive set function on a σ-algebra. Then there exists a partition X = P N into disjoint sets with the following properties. If then if P we must have µ() 0, whereas if N then µ() 0. This partition is essentially unique, in the sense that if X = P N is another such decomposition, then µ (P P ) = 0 and µ (N N ) = 0, where µ will be defined in Definition 8.1.2. 3 See for example [10]. There it is shown that if a series is conditionally convergent, then the sum of the positive terms diverges, and the sum of the negative terms diverges. 133

Proof. Let α = sup{µ() }. Then 0 α < by Theorem 8.1.1 and because µ( ) = 0. For each n N there exists n such that µ( n ) > α 1 2 n. Let P = lim inf n = p=1 n=p so that P and P is the set of all those x X such that x is present in all but a finite number of the sets n. We will show that µ(p) = α. First we need the following lemma. Lemma 8.1.1. Under the hypotheses of Theorem 8.1.2, if µ(b 1 ) > α ǫ 1 and if µ(b 2 ) > α ǫ 2 then µ(b 1 B 2 ) > α (ǫ 1 + ǫ 2 ). Proof. Because µ is (countably) additive, µ(b 1 B 2 ) = µ(b 1 )+µ(b 2 ) µ(b 1 B 2 ) > (α ǫ 1 )+(α ǫ 2 ) α since α = sup{µ() }. = α (ǫ 1 + ǫ 2 ) It follows for each q p that ( q ) µ n > α n=p q n=p n 1 2 α 1 n 2 p 1 for all q p, from which we deduce 4 that ( ) µ n α 1 2 p 1. p 4 The reader should take note that µ need not be monotone, being a signed measure. We use here the boundedness of µ together with its countable additivity to show that ( q ) ( ) µ n µ n as q. n=p 134 n=p

In a similar manner one can show that ( [ ]) ( µ n lim α 1 ) = α. p 2 p 1 p=1 n=p It follows that α µ(p) α, which implies that µ(p) = α as claimed. Moreover, if there were a set such that P and µ() < 0, then we would have µ(p \ ) = µ(p) µ() > µ(p) which is a contradiction. It follows that if P then µ() 0. Now let N = X \P. Suppose there were a measurable set N such that µ() > 0. Then it would follow that µ(p ) > µ(p), which is impossible. Hence if and N it follows that µ() 0. We leave the proof of essential uniqueness to Exercise 8.1.1. Definition 8.1.2. Let µ be a countably additive set function on a σ-algebra, and let P and N be (for µ) as in Theorem 8.1.2. Define the positive part, the negative part, and the variation of µ as follows: µ + () = µ( P) µ () = µ( N) µ () = µ + () + µ () The number µ (X) = µ is called the total variation norm of µ. Exercise 8.1.1. Suppose that we have two Hahn decompositions as in Theorem 8.1.2: X = P N = P N. Prove that µ (P P ) = 0 = µ (N N ). Exercise 8.1.2. (Jordan Decomposition Theorem.) Prove that µ +, µ, and µ are countably additive non-negative measures. Prove also the decomposition µ = µ + µ and that this decomposition is minimal in the following sense. If µ 1 and µ 2 are measures such that µ = µ 1 µ 2, then µ + µ 1 and µ µ 2. Exercise 8.1.3. Prove that the total variation norm satisfies all the requirements to be a norm on the vector space M of all countably additive set functions on (X,, µ). Exercise 8.1.4. Prove that M is complete in the total variation norm. 135

8.2 Radon-Nikodym Theorem Definition 8.2.1. If λ and µ are measures on a σ-algebra of subsets of X, we call λ absolutely continuous with respect to µ, written as λ µ, if and only if and µ() = 0 implies λ() = 0. We have a similar definition for countably additive set functions (signed measures). If a non-negative function f is in L 1 (X,, µ), and if we define λ(e) = f dµ for each E, then λ will be absolutely continuous with respect to µ, written λ µ, as in the foregoing definition. Definition 8.2.2. If λ and µ are countably additive set functions on, we call λ absolutely continuous with respect to µ, written as λ µ, if and only if λ(e) = 0 for each E such that µ (E) = 0. Exercise 8.2.1. Let λ and µ be countably additive set functions. Prove that the following three statements are equivalent. i. λ µ ii. λ + µ and λ µ iii. λ µ Theorem 8.2.1. (Radon-Nikodym) Suppose λ and µ are finite (non-negative) measures on a σ-algebra of subsets of X. Then we have the following conclusions. i. There exists a non-negative function f in L 1 (X,, µ) such that for each we have λ() = f dµ if and only if λ µ. ii. Moreover, the L 1 (X,, µ)-equivalence class of a Radon-Nikodym derivative is uniquely determined. E 136

Remark 8.2.1. For f as in the Radon-Nikodym theorem, it is common to call f the Radon-Nikodym derivative, and to denote this as This notation yields the formula λ() = f = dλ dµ. 1 dλ = dλ dµ dµ which suggests a change of variables formula, and a chain rule. See Exercises 8.2.6 and 8.2.3. Proof of Theorem. The implication from left to right is inherent in the fourth conclusion of Theorem 5.2.2. So we will suppose here that λ µ and give a proof from right to left. We begin with a lemma. Lemma 8.2.1. Under the hypotheses of the Radon-Nikodym theorem, if λ µ, and if λ is not the identically zero measure, then there exists ǫ > 0 and there exists P with µ(p) > 0 such that if and P we have λ() ǫµ(). In the context of this proof, we will write the conclusion of the lemma as an inequality as follows: λ P > ǫµ. Proof of Lemma. Since λ µ and λ is not identically zero, neither is µ identically zero. We claim that there exists sufficiently small ǫ > 0 such that (λ ǫµ) + 0. Suppose this were false. Then we would have λ() ǫµ() for all and for all ǫ > 0. But this would force λ = 0 which is a contradiction. Thus there exists ǫ such that (λ ǫµ) + > 0. Hence there is a Hahn decomposition X = P N for the signed measure (λ ǫµ) such that if a measurable set P implies that (λ ǫµ)() > 0, showing that λ P > ǫµ. 137

If f L + (µ) we define λ f () = f dµ and we let L + (µ, λ) = { f L + (µ) λ f () λ() } noting that the inequalities that define L + (µ, λ) apply to all and are not limited to the subsets of P. By Lemma 8.2.1 we know that there exist ǫ > 0 and P with µ(p) > 0 and such that ǫ1 P L + (µ, λ), which therefore has a non-trivial element if λ is not identically zero. If λ were zero, still the set L + (µ, λ) would be non-empty since it would contain the zero function. Let α = sup{λ f (X) f L + (µ, λ)} so that α λ(x) <. Thus for each n N there exists f n L + (µ, λ) such that λ fn (X) > α 1 n. Let g n = max(f 1,...,f n ) = f 1... f n. Then g n is a monotone increasing sequence of measurable functions, and g n L + (µ, λ) because this is true for each function f j. Moreover, λ gn λ and λ gn (X) > α 1 n. We can define g = lim n g n, which is defined almost everywhere, and λ g λ. We know also that λ g (X) = α. It will suffice to prove that λ g = λ. We will suppose the latter equation is false and deduce a contradiction. Suppose that λ λ g > 0, which means that λ λ g is non-negative and not the identically zero measure. Let λ = λ λ g. Then λ > 0 and λ µ. By Lemma 8.2.1 we conclude that there exists ǫ > 0 and P such that µ(p ) > 0 and such that and P implies that λ() λ g () = λ () ǫµ(). Let h = g + ǫ1 P. Then h dµ = g dµ + ǫµ() for each such that P. That is λ λ g + λ ǫ1p. Hence X h dµ > α, which is impossible since h L+ (µ, λ). The uniqueness of the Radon-Nikodym derivative up to L 1 -equivalence is shown in Exercise 8.2.2. Remark 8.2.2. We remark that if f L 1 (X,, µ) for some measure µ, then the carrier of f must be σ-finite. Thus in order to characterize those 138

measures expressible in the form λ() = f dµ it would be appropriate to limit our attention to σ-finite measure spaces (X,, µ). It is easy to extend the Radon-Nikodym theorem to the case in which µ is a σ-finite measure and λ is a finite measure. It is simple also to give an extension of the Radon-Nikodym theorem to signed measures because each signed measure is the difference between two positive measures and so the Radon-Nikodym derivative in this more general context is the difference between two Radon-Nikodym derivatives for positive measures. Exercise 8.2.2. Show that dλ, the Radon-Nikodym derivative of λ with dµ respect to µ in Theorem 8.2.1, is uniquely determined as an element of L 1 (X,, µ). Exercise 8.2.3. Suppose that the measures λ, µ, ν on a σ-field P(X) have the relationship λ µ ν where λ and µ are finite and ν is σ-finite. Prove that λ ν and that by means of the following steps. a. Let dλ dν = dλ dµ dµ dν f = dλ dµ, g = dµ dν, h = dλ dν and show that there is a sequence f n S 0 such that f n ր f pointwise almost everywhere. b. Show that λ() for all as n. c. Show that as n for all. f n dµ 139 f n dµ 0 f n g dν 0

d. Use Exercise 8.2.2 to complete the proof that f n g n h. Exercise 8.2.4. Let l denote Lebesgue measure on the unit interval, and let φ be the Cantor function from Example 7.4.1. Define λ = l φ by λ() = l(φ()) for each measurable set [0, 1]. Is it true that λ l? Prove your conclusion. Exercise 8.2.5. Let φ be a continuously differentiable monotone increasing function defined on [a, b] R. Define a measure λ on the Lebesgue measurable sets of [a, b] by λ() = l(φ()). Prove that λ l and find dλ dl. Exercise 8.2.6. Suppose (X,, µ) is a complete measure space and f L 1 (X,, µ). Suppose φ : X X is a bijection for which φ(e) if and only if E, and suppose φ maps Lebesgue null sets to Lebesgue null sets. Define the measure µ φ(e) = µ(φ(e)). Prove that µ φ µ and the change of variables formula d(µ φ) (f φ) dµ = f dµ. dµ E Exercise 8.2.7. Suppose µ is a σ-finite measure on a σ-algebra of subsets of X. Suppose λ is another measure on such that λ µ. Prove that there exists a non-negative µ-measurable function f on X such that λ() = f dµ for all. Prove that λ is a finite measure if and only if f L 1 (X,, µ). Exercise 8.2.8. Suppose µ is a σ-finite measure on a σ-algebra of subsets of X. Suppose λ is a signed real-valued measure on such that λ µ. Prove that there exists a (signed) f L 1 (X,, µ) such that λ() = f dµ for all. Exercise 8.2.9. It is interesting to consider the relationship between the concept of absolute continuity of functions given in Definition 7.4.1 and that of absolute continuity of measures. a. If λ and µ are any two finite measures on a σ-field P(X), prove that λ µ if and only if they satisfy the following condition: for each ǫ > 0 there exists a δ > 0 such that µ() < δ implies that λ() < ǫ. (Hint: For one direction, use the Radon-Nikodym theorem.) 140 φ(e)

b. Suppose now that the finite measure λ is defined on the Lebesgue measurable sets of ([a, b), L). Define f(x) = λ[a, x) for all x [a, b]. Prove that f is an absolutely continuous function on [a, b] if and only if λ l. (Hint: From right to left is easy by part (a). For the other direction, prove that λ() = f dl for all L.) 8.3 Lebesgue Decomposition Theorem The Radon-Nikodym Theorem addressed the classification of measures absolutely continuous with respect to a given measure. Here we study a quite different (symmetrical) relationship of singularity between two measures or countably additive set functions. Definition 8.3.1. If λ and µ are countably additive set functions on, we call λ singular with respect to µ if and only if X = E F, a disjoint union of -measurable sets, such that λ (E) = 0 = µ (F). This is denoted as λ µ. Theorem 8.3.1. Let be a σ-algebra of subsets of X. Let µ and ν be two signed measures. Then there exist two unique signed measures ν 0 and ν 1 such that ν = ν 0 + ν 1 with the properties that ν 0 µ and ν 1 µ. Proof. It follows from Definitions 8.2.2 and 8.3.1 that singularity or absolute continuity with respect to a signed measure µ means singularity or absolute continuity with respect to µ. Thus we can assume without loss of generality that µ is a measure. Because of Exercise 8.2.1 and Definition 8.3.1 we can assume without loss of generality that ν is a measure as well 5. The proof of the theorem is based upon the observation that ν (µ+ν). Thus there exists a non-negative measurable function f such that ν(e) = f dµ + f dν E 5 By Definition 8.1.1 our assumption implies that µ and ν are finite measures. For measures that are not signed, the present theorem can be generalized readily to the σ- finite case. See Exercise 8.3.4. 141 E

for all E. Since 0 ν(e) µ(e) + ν(e) we have 0 f 1 ν-almost everywhere and also (µ +ν)-almost everywhere. Let = f 1 (1) and B = f 1 [0, 1). Thus ν() = µ() + ν(). It follows that µ() = 0. Define ν 0 (E) = ν( E) and ν 1 (E) = ν(e B). Thus ν 0 µ because ν 0 vanishes on subsets of c and µ() = 0. Suppose next that µ(e) = 0. Then ν(e B) = 1 dν = f d(µ + ν) = f dν E B E B since µ(e) = 0 by hypothesis. This implies that (1 f) dν = 0. E B Since 1 f 0 ν-almost everywhere, with strict inequality on B, it follows that ν 1 (E) = ν(e B) = 0, so that ν 1 µ. Finally, we prove uniqueness. Let E B ν = ν 0 + ν 1 = ν 0 + ν 1 (8.3) be two Lebesgue decompositions. Thus ν 0 ν 0 = ν 1 ν 1 with one side singular and the other side absolutely continuous with respect to µ. (See Exercise 8.3.1.) This forces both sides to be zero, which completes the proof. (See Exercise 8.3.2.) Exercise 8.3.1. Suppose that the sum of two measures that are absolutely continuous with respect to µ on (X,, µ) must be absolutely continuous. Prove also that the sum of two measures that are singular with respect to µ on (X,, µ) must be singular. Exercise 8.3.2. Let µ and ν be non-negative finite measures on (X, ). If ν µ and ν µ, prove that ν = 0, the identically zero measure on. Exercise 8.3.3. This exercise continues the work begun in Exercise 8.2.9. Let f be a monotone increasing function on [a, b] and define a measure µ by letting it assign to an interval [a, x) the measure µ[a, x) = f(x) f(a). 142

a. Let µ 1 be the absolutely continuous part of µ with respect to Lebesgue measure, and find the Radon-Nikodym derivative dµ 1 dl. b. Show that the singular part µ 0 and the absolutely continuous part µ 1 of µ f can be used to define absolutely continuous and singular parts of the function f. Exercise 8.3.4. Let ν be any σ-finite measure on the measure space (X,, µ), where µ is σ-finite. Prove that there exist two unique measures ν 0 and ν 1 such that ν = ν 0 + ν 1 with the properties that ν 0 µ and ν 1 µ. 143