2.2.2 Monotone Convergence Theorem

Total Page:16

File Type:pdf, Size:1020Kb

2.2.2 Monotone Convergence Theorem CHAPTER 2. THE LEBESGUE INTEGRAL I 22 2.2.2 Monotone convergence theorem The following theorem is what is known in the literature as the "monotone con- vergence theorem" which is a statement about nonnegative increasing sequences (fn)n≥1 of measurable functions in a measure space (Ω; Σ; µ). It has a Corollary which we call here the monotone convergence theorem for summable/integrable 1 functions where the assumption fn ≥ 0 a.e. is replaced by fn 2 L . Contrary to the monotone convergence theorem for the Riemann integral the measurability or integrability of the pointwise (or a.e.) limit is part of the conclusion and not of an hypothesis to be put into the theorem. Monotone convergence theorem - general version: measurable func- + tions. If (fn)n≥1 is a monotone increasing sequence of functions in M (Ω; Σ) which converges almost everywhere to a function f on Ω. Then f is measurable and Z Z Z lim fn(x) dµ = f(x) dµ = lim fn(x) dµ 2 [0; +1]: (2.24) n!1 Ω Ω Ω n!1 R In particular, if lim fn(x) µ(dx) < +1, then f is summable/integrable. n!1 Ω Remark. Note the monotonicity allows us to define f(x) := lim fn(x) 2 [0; +1] n!1 Remark. When the fn are summable/integrable we can drop the assumption that the fn ≥ 0 by considering in that case the non-negative sequence gn = fn − f1. This fact is a corollary of the monotone convergence theorem but we state it as monotone convergence for integrable functions Corollary: Monotone convergence - integrable functions. If (fn)n≥1 is a monotone increasing sequence of functions in L1(Ω; dµ) which converges almost everywhere to a function f on Ω. Then f is measurable and Z Z Z lim fn(x) dµ = f(x) dµ = lim fn(x) dµ 2 [0; +1]: (2.25) n!1 Ω Ω Ω n!1 R In particular, if lim fn(x) µ(dx) < +1, then f is summable/integrable. n!1 Ω Proof of the monotone convergence theorem. First of all, we prove the theorem in the case that fn converges pointwise to f. By the monotonicity of the sequence Sfn (t) ⊂ Sfn+1 (t) for all t and n. By definition 1 [ Sf (t) = Sfn (t) n=1 CHAPTER 2. THE LEBESGUE INTEGRAL I 23 and therefore by the continuity of measures 1 [ lim µ(Sf (t)) = µ Sf (t) = µ(Sf (t)): n!1 n n n=1 note that this convergence is monotone for a sequence of monotone functions. Now the theorem follows from the corresponding theorem of Riemann integra- tion. Indeed, let νn(t) = µ(Sfn (t)) and ν(t) = µ(Sf (t)). Then νn(t) ≤ νn+1(t) ≤ ν(t) for all n 2 N and t > 0 and all functions are decreasing functions of t, hence Riemann integrable on any compact interval. Since obviously Z 1 Z 1 lim νn(t) dt ≤ ν(t) dt n!1 0 0 we have to prove the reversed inequality in the case when the sequence In = R 1 0 νn(t) dt is bounded which implies that νn(t) < 1 and therefore ν(t) < 1 for all t > 0 . Let b > a > 0. Since the νn(t) are decreasing functions of t we have for any positive integer N that N Z b Z b b − a X k(b − a) ν (t) dt ≥ ν (t) dt ≥ ν (a ); a = a + n n N n k k N 0 a k=1 Since ν(t) is a decreasing function of t for any > 0 there is a positive integer N such that N Z b b − a X ν(t) dt ≤ ν(a ) + . N k a k=1 Since νn(t) converges pointwise to ν(t) for all n sufficiently large max jνn(ak) − ν(ak)j < /(b − a) k2f1;:::;Ng with N as selected before. This implies that for all > 0 and all n sufficiently large Z b Z b νn(t) dt ≥ ν(t) dt − 2. a a which implies Z b Z b lim νn(t) dt ≥ ν(t) dt n!1 a a and this inequality proves the claim. Finally we prove the monotone convergence theorem when fn converges to f almost everywhere. We need the following lemma. Lemma 1. Let f 2 M +(Ω; Σ). Then f(x) = 0 almost everywhere if and only if R f dµ = 0. Proof. If R f dµ = 0, then for all a > 0. Z 1 Z a 0 = µ(Sf (t)) dt ≥ µ(Sf (t)) dt ≥ aµ(Sf (a)) 0 0 CHAPTER 2. THE LEBESGUE INTEGRAL I 24 where the latter inequality is called Chebychev's inequality. Hence µ(Sf (a)) = 0 S 1 for all a > 0 Put a = 1=n and note that Sf (0) = Sf ( n ) which has measure n zero thanks to the sigma-additivity of µ. Conversely, let f(x) = 0 almost everywhere. Then µ(Sf (t)) = 0 for all t ≥ 0 and therefore R f dµ = 0 proving the lemma. Proof of the monotone convergence theorem in the case of almost everywhere convergence. Convergence almost everywhere means that there exists a set N ⊂ Ω of measure zero such that fn converges pointwise to f on E := ΩnN. We then apply the monotone convergence theorem to fχE. Indeed, with Lemma 1 we have Z Z Z Z f dµ = fχE dµ = lim fnχE dµ = lim fn dµ. n!1 n!1 2.2.3 Fatou's lemma Fatou's lemma is not only needed to prove the dominated convergence theorem below but it includes also a statement of the behaviour of the integral under pointwise (or a.e.) convergence: The integral is lower semi-continuous under a.e. convergence. + Fatou's lemma. If (fn)n is a sequence of functions in M (Ω; Σ). Then f(x) := lim inffn(x) is measurable and n!1 Z Z Z lim inf fn(x) µ(dx) ≥ f(x) µ(dx) = lim inffn(x) µ(dx) : (2.26) n!1 Ω Ω Ω n!1 In particular, if (fn)n is a sequence of summable functions, then f is summable. Remark. The hypothesis that the fn's are nonnegative is crucial (see exer- cises). Remark. Fatou's lemma means that the mapping f 7! R f is pointwise lower semicontinuous. Corollary. If f(x) = lim fn(x) a.e. , then n!1 Z Z lim inf fn(x) µ(dx) ≥ f(x) µ(dx): n!1 Ω Ω Proof of Fatou's lemma. We have already shown before that f is measur- able. For k ≥ 1 let Fk(x) = infn≥k fn(x). Fk is measurable and if all fn are summable it is also summable since Fk ≤ fk and more generally Fk ≤ fn if n ≥ k. Therefore for all n ≥ k Z Z fn(x) µ(dx) ≥ Fk(x) µ(dx): Ω Ω CHAPTER 2. THE LEBESGUE INTEGRAL I 25 Obviously, Fk is an increasing sequence (i.e Fk+1 ≥ Fk with lim Fk(x) = sup( inf fn(x)) =: lim inffn(x) = supFk(x): k!1 k≥1 n≥k n!1 k≥1 Consequently, applying the monotonicity of the integral and the monotone con- vergence theorem, we have Z Z lim inf fn(x) µ(dx) := sup inf fn(x) µ(dx) n!1 Ω k≥1 n≥k Ω Z Z ≥ lim Fk(x) µ(dx) = f(x) µ(dx): k!1 Ω Ω 2.2.4 Dominated convergence theorem The dominated convergence theorem formulates sufficient conditions under which almost everywhere convergence yields an integrable function and limit and inte- gral are interchangeable. Again this is an important difference between Lebesgue and Riemann integral. Dominated convergence theorem. Let (fn)n be a sequence of (summable) functions in M(Ω; Σ) converging a.e. to a function f. If there exists a nonneg- ative summable function g on (Ω; Σ; µ) such that jfn(x)j ≤ g(x) for all n, then f is summable such that jf(x)j ≤ g(x) and Z Z Z lim fn(x) µ(dx) = f(x) µ(dx) = lim fn(x) µ(dx) : (2.27) n!1 Ω Ω Ω n!1 Proof. Since g − fn ≥ 0 and g + fn ≥ 0 we may apply Fatou's lemma to both sequences of functions: Z Z Z Z g(x)−f(x) µ(dx) ≤ lim inf g(x)−fn(x) µ(dx) ≤ g(x) µ(dx)−lim sup fn(x) µ(dx); Ω n!1 Ω Ω n!1 Ω Z Z Z Z g(x)+f(x) µ(dx) ≤ lim inf g(x)+fn(x) µ(dx) ≤ g(x) µ(dx)+lim inf fn(x) µ(dx): Ω n!1 Ω Ω n!1 Ω Consequently, Z Z Z lim sup fn(x) µ(dx) ≤ f(x) µ(dx) ≤ lim inf fn(x) µ(dx); n!1 Ω Ω n!1 Ω which implies that Z Z lim fn(x) µ(dx) = f(x) µ(dx): n!1 Ω Ω Corollary - convergence in L1. Under the assumptions of the dominated 1 convergence theorem the sequence (fn)n converges in L (Ω; dµ), that is Z lim jfn(x) − f(x)j µ(dx) = 0: (2.28) n!1 Ω CHAPTER 2. THE LEBESGUE INTEGRAL I 26 Proof. Since the sequence (jfn −fj)n also satisfies the hypotheses of the dom- inated convergence theorem (in particular, the jfn − fj are summable since f is summable by the dominated convergence theorem). Now jfn − fj ! 0 a.e.
Recommended publications
  • The Fundamental Theorem of Calculus for Lebesgue Integral
    Divulgaciones Matem´aticasVol. 8 No. 1 (2000), pp. 75{85 The Fundamental Theorem of Calculus for Lebesgue Integral El Teorema Fundamental del C´alculo para la Integral de Lebesgue Di´omedesB´arcenas([email protected]) Departamento de Matem´aticas.Facultad de Ciencias. Universidad de los Andes. M´erida.Venezuela. Abstract In this paper we prove the Theorem announced in the title with- out using Vitali's Covering Lemma and have as a consequence of this approach the equivalence of this theorem with that which states that absolutely continuous functions with zero derivative almost everywhere are constant. We also prove that the decomposition of a bounded vari- ation function is unique up to a constant. Key words and phrases: Radon-Nikodym Theorem, Fundamental Theorem of Calculus, Vitali's covering Lemma. Resumen En este art´ıculose demuestra el Teorema Fundamental del C´alculo para la integral de Lebesgue sin usar el Lema del cubrimiento de Vi- tali, obteni´endosecomo consecuencia que dicho teorema es equivalente al que afirma que toda funci´onabsolutamente continua con derivada igual a cero en casi todo punto es constante. Tambi´ense prueba que la descomposici´onde una funci´onde variaci´onacotada es ´unicaa menos de una constante. Palabras y frases clave: Teorema de Radon-Nikodym, Teorema Fun- damental del C´alculo,Lema del cubrimiento de Vitali. Received: 1999/08/18. Revised: 2000/02/24. Accepted: 2000/03/01. MSC (1991): 26A24, 28A15. Supported by C.D.C.H.T-U.L.A under project C-840-97. 76 Di´omedesB´arcenas 1 Introduction The Fundamental Theorem of Calculus for Lebesgue Integral states that: A function f :[a; b] R is absolutely continuous if and only if it is ! 1 differentiable almost everywhere, its derivative f 0 L [a; b] and, for each t [a; b], 2 2 t f(t) = f(a) + f 0(s)ds: Za This theorem is extremely important in Lebesgue integration Theory and several ways of proving it are found in classical Real Analysis.
    [Show full text]
  • [Math.FA] 3 Dec 1999 Rnfrneter for Theory Transference Introduction 1 Sas Ihnrah U Twl Etetdi Eaaepaper
    Transference in Spaces of Measures Nakhl´eH. Asmar,∗ Stephen J. Montgomery–Smith,† and Sadahiro Saeki‡ 1 Introduction Transference theory for Lp spaces is a powerful tool with many fruitful applications to sin- gular integrals, ergodic theory, and spectral theory of operators [4, 5]. These methods afford a unified approach to many problems in diverse areas, which before were proved by a variety of methods. The purpose of this paper is to bring about a similar approach to spaces of measures. Our main transference result is motivated by the extensions of the classical F.&M. Riesz Theorem due to Bochner [3], Helson-Lowdenslager [10, 11], de Leeuw-Glicksberg [6], Forelli [9], and others. It might seem that these extensions should all be obtainable via transference methods, and indeed, as we will show, these are exemplary illustrations of the scope of our main result. It is not straightforward to extend the classical transference methods of Calder´on, Coif- man and Weiss to spaces of measures. First, their methods make use of averaging techniques and the amenability of the group of representations. The averaging techniques simply do not work with measures, and do not preserve analyticity. Secondly, and most importantly, their techniques require that the representation is strongly continuous. For spaces of mea- sures, this last requirement would be prohibitive, even for the simplest representations such as translations. Instead, we will introduce a much weaker requirement, which we will call ‘sup path attaining’. By working with sup path attaining representations, we are able to prove a new transference principle with interesting applications. For example, we will show how to derive with ease generalizations of Bochner’s theorem and Forelli’s main result.
    [Show full text]
  • Generalizations of the Riemann Integral: an Investigation of the Henstock Integral
    Generalizations of the Riemann Integral: An Investigation of the Henstock Integral Jonathan Wells May 15, 2011 Abstract The Henstock integral, a generalization of the Riemann integral that makes use of the δ-fine tagged partition, is studied. We first consider Lebesgue’s Criterion for Riemann Integrability, which states that a func- tion is Riemann integrable if and only if it is bounded and continuous almost everywhere, before investigating several theoretical shortcomings of the Riemann integral. Despite the inverse relationship between integra- tion and differentiation given by the Fundamental Theorem of Calculus, we find that not every derivative is Riemann integrable. We also find that the strong condition of uniform convergence must be applied to guarantee that the limit of a sequence of Riemann integrable functions remains in- tegrable. However, by slightly altering the way that tagged partitions are formed, we are able to construct a definition for the integral that allows for the integration of a much wider class of functions. We investigate sev- eral properties of this generalized Riemann integral. We also demonstrate that every derivative is Henstock integrable, and that the much looser requirements of the Monotone Convergence Theorem guarantee that the limit of a sequence of Henstock integrable functions is integrable. This paper is written without the use of Lebesgue measure theory. Acknowledgements I would like to thank Professor Patrick Keef and Professor Russell Gordon for their advice and guidance through this project. I would also like to acknowledge Kathryn Barich and Kailey Bolles for their assistance in the editing process. Introduction As the workhorse of modern analysis, the integral is without question one of the most familiar pieces of the calculus sequence.
    [Show full text]
  • Stability in the Almost Everywhere Sense: a Linear Transfer Operator Approach ∗ R
    CORE Metadata, citation and similar papers at core.ac.uk Provided by Elsevier - Publisher Connector J. Math. Anal. Appl. 368 (2010) 144–156 Contents lists available at ScienceDirect Journal of Mathematical Analysis and Applications www.elsevier.com/locate/jmaa Stability in the almost everywhere sense: A linear transfer operator approach ∗ R. Rajaram a, ,U.Vaidyab,M.Fardadc, B. Ganapathysubramanian d a Dept. of Math. Sci., 3300, Lake Rd West, Kent State University, Ashtabula, OH 44004, United States b Dept. of Elec. and Comp. Engineering, Iowa State University, Ames, IA 50011, United States c Dept. of Elec. Engineering and Comp. Sci., Syracuse University, Syracuse, NY 13244, United States d Dept. of Mechanical Engineering, Iowa State University, Ames, IA 50011, United States article info abstract Article history: The problem of almost everywhere stability of a nonlinear autonomous ordinary differential Received 20 April 2009 equation is studied using a linear transfer operator framework. The infinitesimal generator Available online 20 February 2010 of a linear transfer operator (Perron–Frobenius) is used to provide stability conditions of Submitted by H. Zwart an autonomous ordinary differential equation. It is shown that almost everywhere uniform stability of a nonlinear differential equation, is equivalent to the existence of a non-negative Keywords: Almost everywhere stability solution for a steady state advection type linear partial differential equation. We refer Advection equation to this non-negative solution, verifying almost everywhere global stability, as Lyapunov Density function density. A numerical method using finite element techniques is used for the computation of Lyapunov density. © 2010 Elsevier Inc. All rights reserved. 1. Introduction Stability analysis of an ordinary differential equation is one of the most fundamental problems in the theory of dynami- cal systems.
    [Show full text]
  • Chapter 6. Integration §1. Integrals of Nonnegative Functions Let (X, S, Μ
    Chapter 6. Integration §1. Integrals of Nonnegative Functions Let (X, S, µ) be a measure space. We denote by L+ the set of all measurable functions from X to [0, ∞]. + Let φ be a simple function in L . Suppose f(X) = {a1, . , am}. Then φ has the Pm −1 standard representation φ = j=1 ajχEj , where Ej := f ({aj}), j = 1, . , m. We define the integral of φ with respect to µ to be m Z X φ dµ := ajµ(Ej). X j=1 Pm For c ≥ 0, we have cφ = j=1(caj)χEj . It follows that m Z X Z cφ dµ = (caj)µ(Ej) = c φ dµ. X j=1 X Pm Suppose that φ = j=1 ajχEj , where aj ≥ 0 for j = 1, . , m and E1,...,Em are m Pn mutually disjoint measurable sets such that ∪j=1Ej = X. Suppose that ψ = k=1 bkχFk , where bk ≥ 0 for k = 1, . , n and F1,...,Fn are mutually disjoint measurable sets such n that ∪k=1Fk = X. Then we have m n n m X X X X φ = ajχEj ∩Fk and ψ = bkχFk∩Ej . j=1 k=1 k=1 j=1 If φ ≤ ψ, then aj ≤ bk, provided Ej ∩ Fk 6= ∅. Consequently, m m n m n n X X X X X X ajµ(Ej) = ajµ(Ej ∩ Fk) ≤ bkµ(Ej ∩ Fk) = bkµ(Fk). j=1 j=1 k=1 j=1 k=1 k=1 In particular, if φ = ψ, then m n X X ajµ(Ej) = bkµ(Fk). j=1 k=1 In general, if φ ≤ ψ, then m n Z X X Z φ dµ = ajµ(Ej) ≤ bkµ(Fk) = ψ dµ.
    [Show full text]
  • 2. Convergence Theorems
    2. Convergence theorems In this section we analyze the dynamics of integrabilty in the case when se- quences of measurable functions are considered. Roughly speaking, a “convergence theorem” states that integrability is preserved under taking limits. In other words, ∞ if one has a sequence (fn)n=1 of integrable functions, and if f is some kind of a limit of the fn’s, then we would like to conclude that f itself is integrable, as well R R as the equality f = limn→∞ fn. Such results are often employed in two instances: A. When we want to prove that some function f is integrable. In this case ∞ we would look for a sequence (fn)n=1 of integrable approximants for f. B. When we want to construct and integrable function. In this case, we will produce first the approximants, and then we will examine the existence of the limit. The first convergence result, which is somehow primite, but very useful, is the following. Lemma 2.1. Let (X, A, µ) be a finite measure space, let a ∈ (0, ∞) and let fn : X → [0, a], n ≥ 1, be a sequence of measurable functions satisfying (a) f1 ≥ f2 ≥ · · · ≥ 0; (b) limn→∞ fn(x) = 0, ∀ x ∈ X. Then one has the equality Z (1) lim fn dµ = 0. n→∞ X Proof. Let us define, for each ε > 0, and each integer n ≥ 1, the set ε An = {x ∈ X : fn(x) ≥ ε}. ε Obviously, we have An ∈ A, ∀ ε > 0, n ≥ 1. One key fact we are going to use is the following.
    [Show full text]
  • Lecture 26: Dominated Convergence Theorem
    Lecture 26: Dominated Convergence Theorem Continuation of Fatou's Lemma. + + Corollary 0.1. If f 2 L and ffn 2 L : n 2 Ng is any sequence of functions such that fn ! f almost everywhere, then Z Z f 6 lim inf fn: X X Proof. Let fn ! f everywhere in X. That is, lim inf fn(x) = f(x) (= lim sup fn also) for all x 2 X. Then, by Fatou's lemma, Z Z Z f = lim inf fn 6 lim inf fn: X X X If fn 9 f everywhere in X, then let E = fx 2 X : lim inf fn(x) 6= f(x)g. Since fn ! f almost everywhere in X, µ(E) = 0 and Z Z Z Z f = f and fn = fn 8n X X−E X X−E thus making fn 9 f everywhere in X − E. Hence, Z Z Z Z f = f 6 lim inf fn = lim inf fn: X X−E X−E X χ Example 0.2 (Strict inequality). Let Sn = [n; n + 1] ⊂ R and fn = Sn . Then, f = lim inf fn = 0 and Z Z 0 = f dµ < lim inf fn dµ = 1: R R Example 0.3 (Importance of non-negativity). Let Sn = [n; n + 1] ⊂ R and χ fn = − Sn . Then, f = lim inf fn = 0 but Z Z 0 = f dµ > lim inf fn dµ = −1: R R 1 + R Proposition 0.4. If f 2 L and X f dµ < 1, then (a) the set A = fx 2 X : f(x) = 1g is a null set and (b) the set B = fx 2 X : f(x) > 0g is σ-finite.
    [Show full text]
  • MEASURE and INTEGRATION: LECTURE 3 Riemann Integral. If S Is Simple and Measurable Then Sdµ = Αiµ(
    MEASURE AND INTEGRATION: LECTURE 3 Riemann integral. If s is simple and measurable then N � � sdµ = αiµ(Ei), X i=1 �N where s = i=1 αiχEi . If f ≥ 0, then � �� � fdµ = sup sdµ | 0 ≤ s ≤ f, s simple & measurable . X X Recall the Riemann integral of function f on interval [a, b]. Define lower and upper integrals L(f, P ) and U(f, P ), where P is a partition of [a, b]. Set − � � f = sup L(f, P ) and f = inf U(f, P ). P P − A function f is Riemann integrable ⇐⇒ − � � f = f, − in which case this common value is � f. A set B ⊂ R has measure zero if, for any � > 0, there exists a ∞ ∞ countable collection of intervals { Ii}i =1 such that B ⊂ ∪i =1Ii and �∞ i=1 λ(Ii) < �. Examples: finite sets, countable sets. There are also uncountable sets with measure zero. However, any interval does not have measure zero. Theorem 0.1. A function f is Riemann integrable if and only if f is discontinuous on a set of measure zero. A function is said to have a property (e.g., continuous) almost ev­ erywhere (abbreviated a.e.) if the set on which the property does not hold has measure zero. Thus, the statement of the theorem is that f is Riemann integrable if and only if it is continuous almost everywhere. Date: September 11, 2003. 1 2 MEASURE AND INTEGRATION: LECTURE 3 Recall positive measure: a measure function µ: M → [0, ∞] such ∞ �∞ that µ (∪i=1Ei) = i=1 µ(Ei) for Ei ∈ M disjoint.
    [Show full text]
  • Lecture Notes for Math 522 Spring 2012 (Rudin Chapter
    Lebesgue integral We will define a very powerful integral, far better than Riemann in the sense that it will allow us to integrate pretty much every reasonable function and we will also obtain strong convergence results. That is if we take a limit of integrable functions we will get an integrable function and the limit of the integrals will be the integral of the limit under very mild conditions. We will focus only on the real line, although the theory easily extends to more abstract contexts. In Riemann integral the basic block was a rectangle. If we wanted to integrate a function that was identically 1 on an interval [a; b], then the integral was simply the area of that rectangle, so 1×(b−a) = b−a. For Lebesgue integral what we want to do is to replace the interval with a more general subset of the real line. That is, if we have a set S ⊂ R and we take the indicator function or characteristic function χS defined by ( 1 if x 2 S, χ (x) = S 0 else. Then the integral of χS should really be equal to the area under the graph, which should be equal to the \size" of S. Example: Suppose that S is the set of rational numbers between 0 and 1. Let us argue that its size is 0, and so the integral of χS should be 0. Let fx1; x2;:::g = S be an enumeration of the points of S. Now for any > 0 take the sets −j−1 −j−1 Ij = (xj − 2 ; xj + 2 ); then 1 [ S ⊂ Ij: j=1 −j The \size" of any Ij should be 2 , so it seems reasonable to say that the \size" of S is less than the sum of the sizes of the Ij's.
    [Show full text]
  • A Brief Introduction to Lebesgue Theory
    CHAPTER 3 A BRIEF INTRODUCTION TO LEBESGUE THEORY Introduction The span from Newton and Leibniz to Lebesgue covers only 250 years (Figure 3.1). Lebesgue published his dissertation “Integrale,´ longueur, aire” (“Integral, length, area”) in the Annali di Matematica in 1902. Lebesgue developed “measure of a set” in the first chapter and an integral based on his measure in the second. Figure 3.1 From Newton and Leibniz to Lebesgue. Introduction to Real Analysis. By William C. Bauldry 125 Copyright c 2009 John Wiley & Sons, Inc. ⌅ 126 A BRIEF INTRODUCTION TO LEBESGUE THEORY Part of Lebesgue’s motivation were two problems that had arisen with Riemann’s integral. First, there were functions for which the integral of the derivative does not recover the original function and others for which the derivative of the integral is not the original. Second, the integral of the limit of a sequence of functions was not necessarily the limit of the integrals. We’ve seen that uniform convergence allows the interchange of limit and integral, but there are sequences that do not converge uniformly yet the limit of the integrals is equal to the integral of the limit. In Lebesgue’s own words from “Integral, length, area” (as quoted by Hochkirchen (2004, p. 272)), It thus seems to be natural to search for a definition of the integral which makes integration the inverse operation of differentiation in as large a range as possible. Lebesgue was able to combine Darboux’s work on defining the Riemann integral with Borel’s research on the “content” of a set.
    [Show full text]
  • Some Properties of Almost Everywhere Non - Differentiable Functions
    DEMONSTRAHO MATHEMATICA Vol. XVI No 3 1983 Bozena Szkopinska, Janusz Jaskula SOME PROPERTIES OF ALMOST EVERYWHERE NON - DIFFERENTIABLE FUNCTIONS In the paper we shall apply the following notation: Af = {x: | f (x)|<oo} , AfW = {x: f (*) = +00}, Df - the set of all points of discontinuity of a func- tion f, Cf - the set of all points of continuity of the func- tion f, Mf - the set of all those points at which there is no one-sided (finite or infinite) derivative, I - an arbitrary olosed interval with finite Lebesgue measure, R - the Bet of all real numbers, |a| - the Lebesgue measure of a set A, f(x) - the lower derivative of the function f at a point x, f(x) - the upper derivative' of the function f at the point x, A - the closure of the set A. Theorem. If a. set Be Gg, |b| = 0, and there exists a function f s I—R satisfying the conditionsi a) Af(oo) = B b) |l\Mf| = 0, then the set B is nowhere dense. - 695 - 2 B. Szkopinska, J. Jaskuia Proof. It follows from the theorem of Denjoy-Young - -Saks [3] that almost each point x of non-differentiability of a function, in particular, almost each point of the set Mf, is a point at which (1) f(x) = -00 and f(x) = +00. Henoe and from assumption b) it follows that almost each point of the interval I has property (1). This means that the sets {x: f(x) = -00} and {x: f(x) = +00} are dense in I. Whenoe, in virtue of Lemma, [1] p.74, these sets are residual.
    [Show full text]
  • Ultraproducts and the Foundations of Higher Order Fourier Analysis
    Ultraproducts and the Foundations of Higher Order Fourier Analysis Evan Warner Advisor: Nicolas Templier Submitted in partial fulfillment of the requirements for the degree of Bachelor of Arts Department of Mathematics Princeton University May 2012 I hereby declare that this thesis represents my own work in accordance with University regulations. Evan Warner Contents 1. Introduction 1 1.1. Higher order Fourier analysis 1 1.2. Ultraproduct analysis 1 1.3. Overview of the paper 2 2. Preliminaries and a motivating example 4 2.1. Ultraproducts 4 2.2. A motivating example: Szemer´edi’stheorem 11 3. Analysis on ultraproducts 15 3.1. The ultraproduct construction 15 3.2. Measurable functions on the ultraproduct 20 3.3. Integration theory on the ultraproduct 22 4. Product and sub σ algebras 28 − − 4.1. Definitions and the Fubini theorem 28 4.2. Gowers (semi-)norms 31 4.3. Octahedral (semi-)norms and quasirandomness 38 4.4. A few Hilbert space results 40 4.5. Weak orthogonality and coset σ-algebras 42 4.6. Relative separability and essential σ-algebras 45 4.7. Higher order Fourier σ-algebras 48 5. Further results and extensions 52 5.1. The second structure theorem 52 5.2. Prospects for extensions to infinite spaces 58 5.3. Extension to finitely additive measures 60 Appendix A. Proof ofLo´s’ ! theorem and the saturation theorem 62 References 65 1. Introduction 1.1. Higher order Fourier analysis. The subject of higher order Fourier analysis had its genesis in a seminal paper of Gowers, in which the Gowers norms U k for functions on finite groups, where k is a positive integer, were introduced.
    [Show full text]