5 Modes of Convergence

Total Page:16

File Type:pdf, Size:1020Kb

5 Modes of Convergence Tel Aviv University, 2015 Functions of real variables 46 5 Modes of convergence 5a Introduction . 46 5b Local convergence in measure . 47 5c Convergence almost everywhere . 49 5d Dominated convergence; integrals depending on a parameter . 51 5e One-sided bounds . 53 5f Spaces Lp ....................... 56 . 5a Introduction If one has a sequence x1; x2; x3; · · · 2 R of real numbers xn, it is unambiguous what it means for that sequence to converge to a limit x 2 R . If, however, one has a sequence f1; f2; f3;::: of functions fn : X ! R . and a putative limit f : X ! R . there can now be many different ways in which the sequence fn may or may not converge to the limit f . Once X becomes infinite, the functions fn acquire an infinite number of degrees of freedom, and this allows them to approach f in any number of inequivalent ways. However, pointwise and uniform conver- gence are only two of dozens of many other modes of convergence that are of importance in analysis. Tao, Sect. 1.5 \Modes of convergence". When speaking about functions fn on a measure space we really mean, as usual, equivalence classes [fn]. Thus we restrict ourselves to such modes of convergence that are insensitive to arbitrary change of fn on a null set. (Do you bother whether the null set depends on n?) We replace the pointwise convergence with the convergence almost everywhere, fn ! f a.e. () µfx : fn(x) 6! f(x)g = 0 () () fn ! f pointwise outside some null set; and the uniform convergence with the convergence uniform almost every- where, fn ! f in L1 () fn ! f uniformly outside some null set Tel Aviv University, 2015 Functions of real variables 47 (why \L1"? wait for 5f7). On the other hand, the given measure leads to new modes of convergence, such as Z fn ! f in L1 () jfn − fj dµ ! 0 : X 5b Local convergence in measure This is the weakest among all modes of convergence that we need. (That is, every other convergence implies the local convergence in measure.) Strangely enough, it is not mentioned by Tao (Sect. 1.5, p. 95) among 5 modes.1 Throughout, (X; S; µ) is a σ-finite measure space. 5b1 Definition. Let f; f1; f2; ··· : X ! R be measurable functions. We say that fn ! f locally in measure, if µfx 2 A : jfn(x) − f(x)j ≥ "g ! 0 as n ! 1 for all " > 0 and all A 2 S such that µ(A) < 1. This definition generalizes readily to fn : X ! [−∞; +1]; but f must be finite a.e. Clearly, fn ! f if and only if fn − f ! 0. 5b2 Exercise. It is enough to check a single sequence of sets A = A1;A2;::: such that [kAk = X. Prove it.2 If µ(X) < 1, then it is enough to check A = X, of course; this case is called the global convergence in measure. For µ(X) = 1 the global version is stronger than the local one; for example, on R, 1l[n;1) ! 0 locally but not globally. 5b3 Remark. If fn ! f and fn ! g then f = g a.e. (Local convergence in measure is meant in both cases.) Proof (sketch): fx : jf(x) − g(x)j ≥ 2"g ⊂ fx : jfn(x) − f(x)j ≥ "g [ fx : jfn(x) − g(x)j ≥ "g. 1Also not mentioned by Capi´nski& Kopp, nor by Jones. Mentioned by N. Lerner \A course on integration theory", Springer 2014 (Exercise 2.8.14 on p. 113); by F. Liese & K.-J. Miescke \Statistical decision theory", Springer 2008 (Def. A.11 on p. 619); and, prominently, on Wikipedia, \Convergence in measure". 2 Hint: µ A n (A1 [···[ An) ! 0. Tel Aviv University, 2015 Functions of real variables 48 5b4 Remark (sandwich). If gn ≤ fn ≤ hn a.e., gn ! f and hn ! f, then fn ! f. (Local convergence in measure is meant in all three cases.) Proof (sketch): fx : jfn(x) − f(x)j ≥ "g ⊂ fx : jgn(x) − f(x)j ≥ "g [ fx : jhn(x) − f(x)j ≥ "g. 5b5 Exercise. If fn ! 0 and gn ! 0, then afn + bgn ! 0 for arbitrary a; b 2 R. (Local convergence in measure is meant in all three cases; no matter what is meant by 1 − 1.) Prove it. Thus, if fn ! f and gn ! g, then afn + bgn ! af + bg (since (afn + bgn) − (af + bg) = a(fn − f) + b(gn − g)). Interestingly, the local convergence in measure is insensitive not only to the choice of functions fn in their equivalence classes, but also to the choice of a measure µ in its equivalence class defined as follows. 5b6 Definition. A measure ν on (X; S) is called equivalent to µ if ν = f · µ for some f : X ! (0; 1). This is indeed an equivalence relation; symmetry: if ν = f · µ then 1 µ = f · ν; transitivity: if ν1 = f · µ and ν2 = g · ν1 then ν2 = (fg) · µ (recall 4c26 and the paragraph after it). If µ and ν are equivalent then they have the same null sets (think, why). 5b7 Lemma. If ν is equivalent to µ, then fn ! f locally in µ if and only if fn ! f locally in ν. Proof. Let fn ! f locally in µ, " > 0 and ν(A) < 1; we have to prove that ν(An) ! 0, where An = fx 2 A : jfn(x) − f(x)j ≥ "g. By 5b2, WLOG we dν dµ may assume that dµ and dν are bounded on A, since countably many such dµ sets A can cover X (think, why). Now, µ(A) ≤ supx2A dν (x) ν(A) < 1, dν thus µ(An) ! 0, and ν(An) ≤ supx2A dµ (x) µ(An) ! 0. 5b8 Exercise. There exists a finite measure ν equivalent to µ. Prove it.1 For such ν, local convergence in µ is equivalent to the global convergence in ν. The latter is easy to metrize. 5b9 Exercise. If µ(X) < 1, then the following conditions on measurable fn : X ! R are equivalent: (a) fn ! 0 in measure; 1 dν Hint: X = ]Ak, dµ = "k on Ak. Tel Aviv University, 2015 Functions of real variables 49 R (b) X min(1; jfnj) dµ ! 0; (c) R jfnj dµ ! 0. X 1+jfnj Prove it. 1 We may define a metric ρ on the set L0(X) of all equivalence classes of measurable functions X ! R by2 Z (5b10) ρ[f]; [g] = min(1; jf − gj) dµ X provided that µ(X) < 1; otherwise we use a finite equivalent measure in- stead. 5c Convergence almost everywhere As was noted in Sect. 4c, \almost everywhere" (a.e.) means \outside some null set". In particular, we say that fn ! f a.e., if fn(x) ! f(x) for all x outside some null set. (Sometimes this convergence is called \pointwise a.e.".) Once again, the definition generalizes readily to fn : X ! [−∞; +1]; but f must be finite a.e. Once again, fn ! f if and only if fn − f ! 0. Also, if fn ! f and gn ! g, then afn + bgn ! af + bg (no matter what is meant by 1 − 1). Still, (X; S; µ) is a σ-finite measure space. 5c1 Lemma. Convergence almost everywhere implies local convergence in measure. Proof. First, we consider monotone convergence: fn # 0 a.e. In this case, given " > 0 and µ(A) < 1, the sets An = fx 2 A : fn(x) ≥ "g satisfy An #;, that is, A n An " A (up to a null set), which implies µ(A n An) " µ(A), that is, µ(An) # 0. Second, the general case: fn ! 0 a.e. We introduce functions gn = inf(fn; fn+1;::: ), hn = sup(fn; fn+1;::: ); they are measurable by 3c10. Al- 3 most everywhere, gn " 0 (think, why), hn # 0, and gn ≤ fn ≤ hn. By the first part of this proof, gn ! 0 and hn ! 0 locally in measure. It remains to use 5b4. 5c2 Corollary (of 5c1 and 5b3). If fn ! f locally in measure and fn ! g a.e., then f = g a.e. 1 0 Denoted also by L0(X; S; µ), L0(µ), L (X), etc. 2 R jf−gj Alternatively, use X 1+jf−gj dµ. 3Note the zigzag sandwich! Tel Aviv University, 2015 Functions of real variables 50 5c3 Example. Local convergence in measure does not imply convergence almost everywhere. We take the measure space [0; 1] with Lebesgue measure. For every n there exists a partition of X = [0; 1] into n sets (just intervals, if you like) of measure 1=n each. We combine such partitions (for odd n) into a single infinite sequence of sets A1;A2;A3; A4;A5;A6;A7;A8; A9;A10;A11;A12;A13;A14;A15; ::: | {z } | {z } | {z } partition partition partition such that Ak2 ] Ak2+1 ]···] A(k+1)2−1 = X and µ(Ak2 ) = µ(Ak2+1) = 1 ··· = µ(A(k+1)2−1) = 2k+1 for k = 1; 2;::: Clearly, µ(An) ! 0, therefore the indicators fn = 1lAn converge to 0 in measure. However, lim supn fn(x) = 1 for every x, since x belongs to infinitely many An. 5c4 Exercise. If X is (finite or) countable, then local convergence in measure implies convergence almost everywhere. Prove it. P 5c5 Lemma. Let µ(X) < 1, and ρ be defined by (5b10). If n ρ(fn; fn+1) < 1, then the sequence (fn)n converges a.e. Proof. We denote gn = min(1; jfn+1 − fnj), Sn = g1 + ··· + gn " S : X ! [0; 1]. Using the Monotone Convergence Theorem 4c11, Z Z Z Z X Z S dµ = lim Sn dµ = lim g1 dµ+···+ gn dµ = gn dµ < 1 : n n X X X X n X P By (4c31), S < 1 a.e.
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]
  • Modes of Convergence in Probability Theory
    Modes of Convergence in Probability Theory David Mandel November 5, 2015 Below, fix a probability space (Ω; F;P ) on which all random variables fXng and X are defined. All random variables are assumed to take values in R. Propositions marked with \F" denote results that rely on our finite measure space. That is, those marked results may not hold on a non-finite measure space. Since we already know uniform convergence =) pointwise convergence this proof is omitted, but we include a proof that shows pointwise convergence =) almost sure convergence, and hence uniform convergence =) almost sure convergence. The hierarchy we will show is diagrammed in Fig. 1, where some famous theorems that demonstrate the type of convergence are in parentheses: (SLLN) = strong long of large numbers, (WLLN) = weak law of large numbers, (CLT) ^ = central limit theorem. In parameter estimation, θn is said to be a consistent ^ estimator or θ if θn ! θ in probability. For example, by the SLLN, X¯n ! µ a.s., and hence X¯n ! µ in probability. Therefore the sample mean is a consistent estimator of the population mean. Figure 1: Hierarchy of modes of convergence in probability. 1 1 Definitions of Convergence 1.1 Modes from Calculus Definition Xn ! X pointwise if 8! 2 Ω, 8 > 0, 9N 2 N such that 8n ≥ N, jXn(!) − X(!)j < . Definition Xn ! X uniformly if 8 > 0, 9N 2 N such that 8! 2 Ω and 8n ≥ N, jXn(!) − X(!)j < . 1.2 Modes Unique to Measure Theory Definition Xn ! X in probability if 8 > 0, 8δ > 0, 9N 2 N such that 8n ≥ N, P (jXn − Xj ≥ ) < δ: Or, Xn ! X in probability if 8 > 0, lim P (jXn − Xj ≥ 0) = 0: n!1 The explicit epsilon-delta definition of convergence in probability is useful for proving a.s.
    [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]
  • [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]
  • Arxiv:2102.05840V2 [Math.PR]
    SEQUENTIAL CONVERGENCE ON THE SPACE OF BOREL MEASURES LIANGANG MA Abstract We study equivalent descriptions of the vague, weak, setwise and total- variation (TV) convergence of sequences of Borel measures on metrizable and non-metrizable topological spaces in this work. On metrizable spaces, we give some equivalent conditions on the vague convergence of sequences of measures following Kallenberg, and some equivalent conditions on the TV convergence of sequences of measures following Feinberg-Kasyanov-Zgurovsky. There is usually some hierarchy structure on the equivalent descriptions of convergence in different modes, but not always. On non-metrizable spaces, we give examples to show that these conditions are seldom enough to guarantee any convergence of sequences of measures. There are some remarks on the attainability of the TV distance and more modes of sequential convergence at the end of the work. 1. Introduction Let X be a topological space with its Borel σ-algebra B. Consider the collection M˜ (X) of all the Borel measures on (X, B). When we consider the regularity of some mapping f : M˜ (X) → Y with Y being a topological space, some topology or even metric is necessary on the space M˜ (X) of Borel measures. Various notions of topology and metric grow out of arXiv:2102.05840v2 [math.PR] 28 Apr 2021 different situations on the space M˜ (X) in due course to deal with the corresponding concerns of regularity. In those topology and metric endowed on M˜ (X), it has been recognized that the vague, weak, setwise topology as well as the total-variation (TV) metric are highlighted notions on the topological and metric description of M˜ (X) in various circumstances, refer to [Kal, GR, Wul].
    [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]
  • Noncommutative Ergodic Theorems for Connected Amenable Groups 3
    NONCOMMUTATIVE ERGODIC THEOREMS FOR CONNECTED AMENABLE GROUPS MU SUN Abstract. This paper is devoted to the study of noncommutative ergodic theorems for con- nected amenable locally compact groups. For a dynamical system (M,τ,G,σ), where (M, τ) is a von Neumann algebra with a normal faithful finite trace and (G, σ) is a connected amenable locally compact group with a well defined representation on M, we try to find the largest non- commutative function spaces constructed from M on which the individual ergodic theorems hold. By using the Emerson-Greenleaf’s structure theorem, we transfer the key question to proving the ergodic theorems for Rd group actions. Splitting the Rd actions problem in two cases accord- ing to different multi-parameter convergence types—cube convergence and unrestricted conver- gence, we can give maximal ergodic inequalities on L1(M) and on noncommutative Orlicz space 2(d−1) L1 log L(M), each of which is deduced from the result already known in discrete case. Fi- 2(d−1) nally we give the individual ergodic theorems for G acting on L1(M) and on L1 log L(M), where the ergodic averages are taken along certain sequences of measurable subsets of G. 1. Introduction The study of ergodic theorems is an old branch of dynamical system theory which was started in 1931 by von Neumann and Birkhoff, having its origins in statistical mechanics. While new applications to mathematical physics continued to come in, the theory soon earned its own rights as an important chapter in functional analysis and probability. In the classical situation the sta- tionarity is described by a measure preserving transformation T , and one considers averages taken along a sequence f, f ◦ T, f ◦ T 2,..
    [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]
  • Sequences and Series of Functions, Convergence, Power Series
    6: SEQUENCES AND SERIES OF FUNCTIONS, CONVERGENCE STEVEN HEILMAN Contents 1. Review 1 2. Sequences of Functions 2 3. Uniform Convergence and Continuity 3 4. Series of Functions and the Weierstrass M-test 5 5. Uniform Convergence and Integration 6 6. Uniform Convergence and Differentiation 7 7. Uniform Approximation by Polynomials 9 8. Power Series 10 9. The Exponential and Logarithm 15 10. Trigonometric Functions 17 11. Appendix: Notation 20 1. Review Remark 1.1. From now on, unless otherwise specified, Rn refers to Euclidean space Rn n with n ≥ 1 a positive integer, and where we use the metric d`2 on R . In particular, R refers to the metric space R equipped with the metric d(x; y) = jx − yj. (j) 1 Proposition 1.2. Let (X; d) be a metric space. Let (x )j=k be a sequence of elements of X. 0 (j) 1 Let x; x be elements of X. Assume that the sequence (x )j=k converges to x with respect to (j) 1 0 0 d. Assume also that the sequence (x )j=k converges to x with respect to d. Then x = x . Proposition 1.3. Let a < b be real numbers, and let f :[a; b] ! R be a function which is both continuous and strictly monotone increasing. Then f is a bijection from [a; b] to [f(a); f(b)], and the inverse function f −1 :[f(a); f(b)] ! [a; b] is also continuous and strictly monotone increasing. Theorem 1.4 (Inverse Function Theorem). Let X; Y be subsets of R.
    [Show full text]
  • Uniqueness for the Signature of a Path of Bounded Variation and the Reduced Path Group
    ANNALS OF MATHEMATICS Uniqueness for the signature of a path of bounded variation and the reduced path group By Ben Hambly and Terry Lyons SECOND SERIES, VOL. 171, NO. 1 January, 2010 anmaah Annals of Mathematics, 171 (2010), 109–167 Uniqueness for the signature of a path of bounded variation and the reduced path group By BEN HAMBLY and TERRY LYONS Abstract We introduce the notions of tree-like path and tree-like equivalence between paths and prove that the latter is an equivalence relation for paths of finite length. We show that the equivalence classes form a group with some similarity to a free group, and that in each class there is a unique path that is tree reduced. The set of these paths is the Reduced Path Group. It is a continuous analogue of the group of reduced words. The signature of the path is a power series whose coefficients are certain tensor valued definite iterated integrals of the path. We identify the paths with trivial signature as the tree-like paths, and prove that two paths are in tree-like equivalence if and only if they have the same signature. In this way, we extend Chen’s theorems on the uniqueness of the sequence of iterated integrals associated with a piecewise regular path to finite length paths and identify the appropriate ex- tended meaning for parametrisation in the general setting. It is suggestive to think of this result as a noncommutative analogue of the result that integrable functions on the circle are determined, up to Lebesgue null sets, by their Fourier coefficients.
    [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]