FUNDAMENTALS of REAL ANALYSIS by Do˘Gan C¸Ömez IV. DIFFERENTIATION and SIGNED MEASURES IV.1. Differentiation of Monotonic

Total Page:16

File Type:pdf, Size:1020Kb

FUNDAMENTALS of REAL ANALYSIS by Do˘Gan C¸Ömez IV. DIFFERENTIATION and SIGNED MEASURES IV.1. Differentiation of Monotonic FUNDAMENTALS OF REAL ANALYSIS by Do˘ganC¸¨omez IV. DIFFERENTIATION AND SIGNED MEASURES IV.1. Differentiation of Monotonic Functions R Question: Can we calculate E f easily? More explicitly, can we hope for a result like the Fundamental Theorem of Calculus for Riemann Integral? Answer: Yes, but with lots of work and patience. The following fact is well-known: Proposition. Let f :[a; b] ! R be a monotonic function. Then the set of discontinuities of f is at most countable (hence has measure 0). Exercise. Prove that if E be a countable subset of [a; b], then there exists f :(a; b) ! R which is continuous only on (a; b) n E. Convention: Throughout this chapter, [a; b] will be an interval with −∞ < a < b < 1. Definition. A collection F of closed and bounded intervals is called a Vitali cover for a set E ⊂ R if for all x 2 E, for all > 0 there exists I 2 F such that x 2 I and `(I) < . Example. Let E 2 R be a bounded set (say E ⊂ [c; d]). Then the collection F = f[α; β]: c ≤ α < β ≤ d; α; β 2 Qg is a Vitali Cover for E. Next theorem is an important one! Theorem.1 (Vitali Covering Lemma) Let E ⊂ R with m∗(E) < 1 and let F be a Vitali n cover for E. Then for all > 0 there exists a finite disjoint sub-collection fIkgk=1 ⊂ F such that n ! ∗ [ m E n Ik < . k=1 Comments: (a) One can always find an open set O such that E ⊂ O and m(O) < 1. (b) Without loss of generality, we can assume that [ X I 2 F ) I ⊂ O ) m Ik = m(Ik) ≤ m(O) < 1: | {z } =`(Ik) Hence `(Ik) ! 0 if the collection F is infinite. (c) For any interval I, Ie will be the closed interval having the same midpoint as I and has length 5 times the length of I. Proof (of Theorem.1) We will consider two cases. n Case 1: There exists a disjoint sub-collection fIkgk=1 ⊂ F such that for all I 2 F, I \ Sn ( k=1 Ik) 6= ;. 1 2 Sn In this case we claim that E ⊂ k=1 Ik (and hence the theorem is proven). If the claim is Sn false, then there exists a point x 2 E n k=1 Ik. Then x 62 Ik for all 1 ≤ k ≤ n. Since F is a Vitali cover, there exists I 2 F such that x 2 I. By assumption I \ Ik0 6= ; for some 1 ≤ k0 ≤ k. Without loss of generality we can assume that I \ Ik = ; for k 6= k0. First 1 of all I 6= Ik0 , hence `(I n Ik0 ) > 0. Let < 2 `(I n Ik), then there exists J 2 F such that x 2 J and `(J) < . Thus n ! [ J \ Ik = ;: k=1 This contradicts our assumption, proving the claim. n Case 2: No such finite family exists. That is, given any finite disjoint collection fIkgk=1; there Sn exists I 2 F such that I \ ( k=1 Ik) = ;. 1 [In this case we will construct (inductively) a countably disjoint sub-collection fIkg 2 F S k=1 such that E n k Ik is as small as we wish.] Let I1 be arbitrary. By assumption there exists I (there may be others) such that I \ I1 = ;. 1 Pick I1 2 F such that I2 \ I1 = ; and `(I2) > 2 S1 where S1 = `(I). Now assume we have a n disjoint collection fIkgk=1 ⊂ F constructed in this manner. Then let Sn = supf`(I): I 2 F;I \ Ik = ;; k = 1; : : : ng: 1 n+1 Pick In+1 2 F such that In+1 \ Ik = ; for all k = 1; : : : n and `(In+1) > 2 Sn. So fIkgk=1 is a 1 disjoint family. Hence, by induction we have a countably disjoint family fIkgk=1 ⊂ F such that each Ik ⊂ O for all k; where O is an open set such that E ⊂ O and m(O) < 1. Then `(Ik) ! 0 must hold. Claim: This collection has the property that for all n ≥ 1 n 1 [ [ E n Ik ⊂ Iek: k=1 k=n+1 Sn To prove this claim let n ≥ 1 and let x 2 E n k=1 Ik. Then there exists I 2 F such that x 2 I and n ! [ I \ Ik = ;: k=1 Then I \ Im 6= ; for some m ≥ n + 1 must hold (otherwise, if I \ Im = ; for all m, then 1 1 `(Im) > 2 `(I) for all m, which implies that `(Em) 6! 0). So now we have `(Im) ≥ 2 Sm. Let cm be the center of Im. So 1 1 jx − c j < `(I ) + `(I) < `(I ) + 2`(I ): m 2 m 2 n m S1 Hence x must belong to Iem. Therefore x 2 k=n+1 Iek. This implies n 1 [ [ E n Ik ⊂ Iek: k=1 k=n+1 This now proves the above claim. 3 S1 P1 So given > 0, since fIkg are disjoint, m( k=1 Ik) = k=1 `(Ik) < m(O). Hence, we can pick P1 n such that k=n+1 `(Ik) < 5 . Then n ! 1 ! 1 ∗ [ ∗ [ X m E n Ik ≤ m Iek ≤ `(Iek) < . k=1 k=n+1 k=n+1 Observe: if f :[a; b] ! R is continuous and differentiable on (a; b), by the Mean Value Theorem, if f 0(x) > α > 0 on (c; d) ⊂ (a; b), then f(d) − f(c) ≥ α(d − c): (∗) Exercise.1 Let f :[a; b] ! R be a continuous function which is differentiable on (a; b): Let E = fx 2 (a; b): f 0(x) > αg and let F = f[c; d] ⊂ (a; b) : (*) holdsg: Then the collection F is a Vitali Cover for E. Definition. Let f :[a; b] ! R. The upper derivative of f at c 2 (a; b) is the value (possibly +1) defined by " # f(x + t) − f(x) Df(c) = lim sup : h!0 0<jt|≤h t Similarly, the lower derivative of f at c 2 (a; b) is f(x + t) − f(x) Df(c) = lim inf : h!0 0<jt|≤h t If Df(c) = Df(c) and is < 1, then we say that f is differentiable at c, and call this common value as the derivative of f at c; which is denoted by f 0(c). Exercise.2 (i) Let f :[−1; 1] ! R be given as 8 < x if 0 ≤ x 2 Q f(x) = 0 if 0 ≤ x 2 R n Q : x 2 if x < 0: Find Df(0) and Df(0). (ii) Construct f :[a; b] ! R continuous such that Df(c) 6= Df(c); for some c 2 (a; b): (iii) Construct f :[a; b] ! R such that Df(c) = 1 whereas Df(c) = 0; for some c 2 (a; b): Fact.1 Let f :[a; b] ! R be increasing. Then for all α > 0 1 (a) m∗(fx 2 (a; b): Df(x) ≥ αg) ≤ [f(b) − f(a)] and α (b) m∗(fx 2 (a; b): Df(x) = +1g) = 0: Proof. (a) Let α; > 0 be arbitrary. Let Eα = fx 2 (a; b): Df(x) ≥ αg: Choose 0 < β < α, and let F = f[c; d] ⊂ (a; b): f(d) − f(c) ≥ β(d − c)g: Then F is a Vitali Cover for Eα. Then by the Vitali Covering Lemma, " n # " n #! n ! ∗ ∗ [ [ ∗ [ m (Eα) = m Eα n [ck; dk] [ [ck; dk] < + m [ck; dk] k=1 k=1 k=1 4 n for some disjoint sub-collection f[ck; dk]gk=1 ⊂ F. Hence n ! n n [ X 1 X + m∗ [c ; d ] ≤ + (d − c ) < + [f(d ) − f(c )] k k k k β k k k=1 k=1 k=1 1 1 ≤ + [f(b) − f(a)] ) m∗(E ) < [f(b) − f(a)]: β α α 1 1 For E = fx 2 (a; b): Df(x) = +1g, E ⊂ EN for any N ≥ 1. Therefore, 1 m∗(E1) ≤ m∗(E ) ≤ [f(d) − f(c)] −! 0 as N ! 1; implying (b): N N Corollary. If f is monotonic and differentiable, then f 0 < 1 almost everywhere. Question: Which functions are differentiable? (In particular, are monotonic functions differen- tiable?) Theorem.2 (Lebesgue Differentiation Theorem) Let f :(a; b) ! R be monotonic. Then f is differentiable almost everywhere on (a; b). Proof. We will assume that f is increasing on (a; b). Let Eα,β = fx 2 (a; b): Df(x) > β > α > Df(x)g: Note that f fails to be differentiable on points at which Df(x) < Df(x). Then the family fEα,βgα,β2Q is a countable collection of sets on which f is not differentiable. Hence, if [ E = Eα,β; α,β2Q α<β ∗ ∗ we need to show that m (E) = 0, which will be the case if we show that m (Eα,β) = 0 for each α and β: So, fix α < β and let F = Eα,β. Given > 0; pick O open such that F ⊂ O ⊂ (a; b) and m(O) < m∗(F ) + . Let F = f[c; d]: f(d) − f(c) < α(d − c)g: Then F is a Vitali Cover for F by Exercise.1. Hence by the Vitali Covering Lemma there exists n a finite collection f[ck; dk]gk=1 ⊂ F such that n ! ∗ [ m F n [ck; dk] < . k=1 Then n n n ! X X [ [f(dk) − f(ck)] < α (dk − ck) = αm [ck; dk] k=1 k=1 k=1 ≤ αm(O) < α(m∗(F ) + ): Now focus on the interval (ck; dk).
Recommended publications
  • Lectures on Fractal Geometry and Dynamics
    Lectures on fractal geometry and dynamics Michael Hochman∗ March 25, 2012 Contents 1 Introduction 2 2 Preliminaries 3 3 Dimension 3 3.1 A family of examples: Middle-α Cantor sets . 3 3.2 Minkowski dimension . 4 3.3 Hausdorff dimension . 9 4 Using measures to compute dimension 13 4.1 The mass distribution principle . 13 4.2 Billingsley's lemma . 15 4.3 Frostman's lemma . 18 4.4 Product sets . 22 5 Iterated function systems 24 5.1 The Hausdorff metric . 24 5.2 Iterated function systems . 26 5.3 Self-similar sets . 31 5.4 Self-affine sets . 36 6 Geometry of measures 39 6.1 The Besicovitch covering theorem . 39 6.2 Density and differentiation theorems . 45 6.3 Dimension of a measure at a point . 48 6.4 Upper and lower dimension of measures . 50 ∗Send comments to [email protected] 1 6.5 Hausdorff measures and their densities . 52 1 Introduction Fractal geometry and its sibling, geometric measure theory, are branches of analysis which study the structure of \irregular" sets and measures in metric spaces, primarily d R . The distinction between regular and irregular sets is not a precise one but informally, k regular sets might be understood as smooth sub-manifolds of R , or perhaps Lipschitz graphs, or countable unions of the above; whereas irregular sets include just about 1 everything else, from the middle- 3 Cantor set (still highly structured) to arbitrary d Cantor sets (irregular, but topologically the same) to truly arbitrary subsets of R . d For concreteness, let us compare smooth sub-manifolds and Cantor subsets of R .
    [Show full text]
  • 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]
  • Measure and Integration (Under Construction)
    Measure and Integration (under construction) Gustav Holzegel∗ April 24, 2017 Abstract These notes accompany the lecture course ”Measure and Integration” at Imperial College London (Autumn 2016). They follow very closely the text “Real-Analysis” by Stein-Shakarchi, in fact most proofs are simple rephrasings of the proofs presented in the aforementioned book. Contents 1 Motivation 3 1.1 Quick review of the Riemann integral . 3 1.2 Drawbacks of the class R, motivation of the Lebesgue theory . 4 1.2.1 Limits of functions . 5 1.2.2 Lengthofcurves ....................... 5 1.2.3 TheFundamentalTheoremofCalculus . 5 1.3 Measures of sets in R ......................... 6 1.4 LiteratureandFurtherReading . 6 2 Measure Theory: Lebesgue Measure in Rd 7 2.1 Preliminaries and Notation . 7 2.2 VolumeofRectanglesandCubes . 7 2.3 Theexteriormeasure......................... 9 2.3.1 Examples ........................... 9 2.3.2 Propertiesoftheexteriormeasure . 10 2.4 Theclassof(Lebesgue)measurablesets . 12 2.4.1 The property of countable additivity . 14 2.4.2 Regularity properties of the Lebesgue measure . 15 2.4.3 Invariance properties of the Lebesgue measure . 16 2.4.4 σ-algebrasandBorelsets . 16 2.5 Constructionofanon-measurableset. 17 2.6 MeasurableFunctions ........................ 19 2.6.1 Some abstract preliminary remarks . 19 2.6.2 Definitions and equivalent formulations of measurability . 19 2.6.3 Properties 1: Behaviour under compositions . 21 2.6.4 Properties 2: Behaviour under limits . 21 2.6.5 Properties3: Behaviourof sums and products . 22 ∗Imperial College London, Department of Mathematics, South Kensington Campus, Lon- don SW7 2AZ, United Kingdom. 1 2.6.6 Thenotionof“almosteverywhere”. 22 2.7 Building blocks of integration theory .
    [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]
  • Some Integral Inequalities for Operator Monotonic Functions on Hilbert Spaces Received May 6, 2020; Accepted June 24, 2020
    Spec. Matrices 2020; 8:172–180 Research Article Open Access Silvestru Sever Dragomir* Some integral inequalities for operator monotonic functions on Hilbert spaces https://doi.org/10.1515/spma-2020-0108 Received May 6, 2020; accepted June 24, 2020 Abstract: Let f be an operator monotonic function on I and A, B 2 SAI (H) , the class of all selfadjoint op- erators with spectra in I. Assume that p : [0, 1] ! R is non-decreasing on [0, 1]. In this paper we obtained, among others, that for A ≤ B and f an operator monotonic function on I, Z1 Z1 Z1 0 ≤ p (t) f ((1 − t) A + tB) dt − p (t) dt f ((1 − t) A + tB) dt 0 0 0 1 ≤ [p (1) − p (0)] [f (B) − f (A)] 4 in the operator order. Several other similar inequalities for either p or f is dierentiable, are also provided. Applications for power function and logarithm are given as well. Keywords: Operator monotonic functions, Integral inequalities, Čebyšev inequality, Grüss inequality, Os- trowski inequality MSC: 47A63, 26D15, 26D10. 1 Introduction Consider a complex Hilbert space (H, h·, ·i). An operator T is said to be positive (denoted by T ≥ 0) if hTx, xi ≥ 0 for all x 2 H and also an operator T is said to be strictly positive (denoted by T > 0) if T is positive and invertible. A real valued continuous function f (t) on (0, ∞) is said to be operator monotone if f (A) ≥ f (B) holds for any A ≥ B > 0. In 1934, K. Löwner [7] had given a denitive characterization of operator monotone functions as follows: Theorem 1.
    [Show full text]
  • Dirichlet Is Natural Vincent Danos, Ilias Garnier
    Dirichlet is Natural Vincent Danos, Ilias Garnier To cite this version: Vincent Danos, Ilias Garnier. Dirichlet is Natural. MFPS 31 - Mathematical Foundations of Program- ming Semantics XXXI, Jun 2015, Nijmegen, Netherlands. pp.137-164, 10.1016/j.entcs.2015.12.010. hal-01256903 HAL Id: hal-01256903 https://hal.archives-ouvertes.fr/hal-01256903 Submitted on 15 Jan 2016 HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est archive for the deposit and dissemination of sci- destinée au dépôt et à la diffusion de documents entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non, lished or not. The documents may come from émanant des établissements d’enseignement et de teaching and research institutions in France or recherche français ou étrangers, des laboratoires abroad, or from public or private research centers. publics ou privés. MFPS 2015 Dirichlet is natural Vincent Danos1 D´epartement d'Informatique Ecole normale sup´erieure Paris, France Ilias Garnier2 School of Informatics University of Edinburgh Edinburgh, United Kingdom Abstract Giry and Lawvere's categorical treatment of probabilities, based on the probabilistic monad G, offer an elegant and hitherto unexploited treatment of higher-order probabilities. The goal of this paper is to follow this formulation to reconstruct a family of higher-order probabilities known as the Dirichlet process. This family is widely used in non-parametric Bayesian learning. Given a Polish space X, we build a family of higher-order probabilities in G(G(X)) indexed by M ∗(X) the set of non-zero finite measures over X. The construction relies on two ingredients.
    [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]
  • Pathological Real-Valued Continuous Functions
    Pathological Real-Valued Continuous Functions by Timothy Miao A project submitted to the Department of Mathematical Sciences in conformity with the requirements for Math 4301 (Honour's Seminar) Lakehead University Thunder Bay, Ontario, Canada copyright c (2013) Timothy Miao Abstract We look at continuous functions with pathological properties, in particular, two ex- amples of continuous functions that are nowhere differentiable. The first example was discovered by K. W. T. Weierstrass in 1872 and the second by B. L. Van der Waerden in 1930. We also present an example of a continuous strictly monotonic function with a vanishing derivative almost everywhere, discovered by Zaanen and Luxemburg in 1963. i Acknowledgements I would like to thank Razvan Anisca, my supervisor for this project, and Adam Van Tuyl, the course coordinator for Math 4301. Without their guidance, this project would not have been possible. As well, all the people I have learned from and worked with from the Department of Mathematical Sciences at Lakehead University have contributed to getting me to where I am today. Finally, I am extremely grateful for the support that my family has given during all of my education. ii Contents Abstract i Acknowledgements ii Chapter 1. Introduction 1 Chapter 2. Continuity and Differentiability 2 1. Preliminaries 2 2. Sequences of functions 4 Chapter 3. Two Continuous Functions that are Nowhere Differentiable 6 1. Example given by Weierstrass (1872) 6 2. Example given by Van der Waerden (1930) 10 Chapter 4. Monotonic Functions and their Derivatives 13 1. Preliminaries 13 2. Some notable theorems 14 Chapter 5. A Strictly Monotone Function with a Vanishing Derivative Almost Everywhere 18 1.
    [Show full text]
  • The Fixed-Point Theorem 8 March 2013 Lecturer: Andrew Myers
    CS 6110 Lecture 21 The Fixed-Point Theorem 8 March 2013 Lecturer: Andrew Myers We saw that the semantics of the while command are a fixed point. We also saw that intuitively, the semantics are the limit of a series of approximations capturing a finite number of iterations of the loop, and giving a result of ? for greater numbers of iterations. In order to take a limit, we need greater structure, which led us to define partial orders. But ordering is not enough. 1 Complete partial orders (CPOs) Least upper bounds Given a partial order (S; v), and a subset B ⊆ S, y is an upper bound of B iff 8x 2 B:x v y. In addition, y is a least upper bound iff y is an upper bound and y v z for all upper bounds z of B. We may abbreviate “least upper bound” as LUB or lub. We notate the LUB of a subset B as F B. We may also make this an infix operator, F F writing i21::m xi = x1 t ::: t xm = fxigi21::m. This is also known as the join of elements x1; : : : ; xm. Chains A chain is a pairwise comparable sequence of elements from a partial order (i.e., elements x0; x1; x2 ::: F such that x0 v x1 v x2 v :::). For any finite chain, its LUB is its last element (e.g., xi = xn). Infinite chains (!-chains, i.e. indexed by the natural numbers) may also have LUBs. Complete partial orders A complete partial order (CPO)1 is a partial order in which every chain has a least upper bound.
    [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]
  • Fixpoints and Applications
    H250: Honors Colloquium – Introduction to Computation Fixpoints and Applications Marius Minea [email protected] Fixpoint Definition Given f : X → X, a fixpoint (fixed point) of f is a value c with f (c) = c. A function may have 0, 1, or many fixpoints. Examples: fixpoint for a real-valued function: intersection with y = x fixpoint for a permutation of 1 .. n Think of the function as a transformation (which may be repeated) Fixpoint: no change Intuitive Examples All-pairs shortest paths in graph Simpler: Path relation in graph Questions: When do such computations terminate? How to express this with fixpoints? Partially Ordered Sets Recall: partial order on a set: reflexive antisymmetric transitive A set A together with a partial order v on A is called a partially ordered set (poset) hA, vi Lattices Many familiar partial orders have additional properties: - any two elements have a unique least upper bound (join t) least x with a v x and b v x - any two elements have a unique greatest lower bound (meet u) A poset with these properties is called a lattice. Inductively: any finite set has a least upper / greatest lower bound. Complete lattice: any subset has least upper/greatest lower bound. =⇒ lattice has a top > and bottom ⊥ element. Iterating a function Define f n(x) = f ◦ f ◦ ... ◦ f (x). | {z } n times On a finite set, iteration will - close a cycle, or - reach a fixpoint (particular case) For an infinite set, iteration may be infinite (none of the above) Monotonic Functions Given a poset hS, vi, a function f : S → S is monotonic if it is either - increasing, ∀x∀y : x v y → f (x) v f (y) - decreasing, ∀x∀y : x v y → f (x) w f (y) Knaster-Tarski Fixpoint Theorem A monotonic function on a complete lattice has a least fixpoint and a greatest fixpoint.
    [Show full text]