Math 641 Lecture #45 ¶9.5,9.6 the Codomain of the Fourier Transform

Total Page:16

File Type:pdf, Size:1020Kb

Math 641 Lecture #45 ¶9.5,9.6 the Codomain of the Fourier Transform Math 641 Lecture #45 ¶9.5,9.6 The Codomain of the Fourier Transform Question. Into what space does the Fourier transform send L1(R) functions? Definition. Let f be any function on R. For each y ∈ R, the translate of f by y is the function fy defined by fy(x) = f(x − y). p Theorem (9.5). Let 1 ≤ p < ∞. For each f ∈ L (R), the mapping y → fy from R into Lp(R) is uniformly continuous. Proof. Let f ∈ Lp(R), and fix > 0. p Since Cc(R) is dense in L (R), there is a g ∈ Cc(R) such that supp(g) ⊂ [−A, A] for A > 0, and kf − gkp < . Since a continuous function with a compact support is uniformly continuous, there is δ ∈ (0,A)[δ = min{δ∗, A/2} where δ∗ corresponds to ∗] such that |s − t| < δ ⇒ |g(s) − g(t)| < (3A)−1/p [(3A)−1/p = ∗]. For any s, t ∈ R with |s − t| < δ, it follows for any x ∈ R that |(x − s) − (x − t)| < δ. Since supp(g) ⊂ [−A, A], the quantity |g(x − s) − g(x − t)| is zero when |x − s| ≥ A and |x − t| ≥ A. Only when |x − s| < A or |x − t| < A, is it possible for this quantity to be nonzero. This requires that −A + s < x < A + s or −A + t < x < A + t, which is min{−A + s, −A + t} < x < max{A + s, A + t}. The difference max{A + s, A + t} − min{−A + s, −A + t} is at most 2A + δ. Thus Z ∞ Z ∞ p p |gs(x) − gt(x)| dx = |g(x − s) − g(x − t)| dx −∞ −∞ Z max{A+s,A+t} = |g(x − s) − g(x − t)|p dx < (3A)−1p(2A + δ) < p, min{−A+s,−A+t} so that kgs − gtkp < . Since the Lp-norms are translation invariant (with respect to Lesbegue measure), kfs − ftkp = kfs − gskp + kgs − gtkp + kgt − ftkp = k(f − g)skp + kgs − gtkp + k(f − g)tkp < 3 whenever |s − t| < δ. Therefore y → fy is uniformly continuous on R. 1 ˆ ˆ Theorem (9.6). If f ∈ L (R), then f ∈ C0(R) and kfku ≤ kfk1. Proof. The Fourier transform of f ∈ L1(R) is in L∞(R): Z ∞ Z ∞ ˆ −ixt |f(t)| = f(x)e dm(x) ≤ |f(x)| dm(x) = kfk1, −∞ −∞ ˆ so that kfk∞ ≤ kfk1. ∞ ˆ ˆ ˆ Recall that C0(R) ⊂ L (R) where kfku = kfk∞ when f ∈ C0(R). It remains to show that f ∈ C0(R). −itnx −itx Fix t ∈ R, let tn → t, and set gn(x) = |f(x)| |e − e |. 1 The sequence {gn} converges pointwise to 0, and is dominated by 2|f| ∈ L (R). By LDCT, Z ∞ ˆ ˆ −itnx itx lim |f(tn) − f(t)| ≤ lim |f(x)| |e − e | dm(x) n→∞ n→∞ −∞ Z ∞ = lim gn(x) dm(x) = 0, −∞ n→∞ so that fˆ is continuous. Now consider t with |t| large. Since eiπ = −1, then Z ∞ fˆ(t) = f(x)e−itxe−iπeiπ dm(x) −∞ Z ∞ = eiπ f(x)e−it(x+π/t) dm(x) −∞ Z ∞ = − f(x − π/t)e−itxdm(x). −∞ This implies that 2fˆ(t) = fˆ(t) + fˆ(t) Z ∞ Z ∞ = f(x)e−ixt dm(x) − f(x − t/π)e−itx dm(x) −∞ −∞ Z ∞ π = f(x) − f x − e−ixt dm(x) −∞ t so that ˆ 2|f(t)| ≤ kf0 − fπ/tk1. ˆ By Theorem 9.5, kf0 − fπ/tk1 → 0 as |0 − π/t| → 0, so that |f(t)| → 0 as |t| → ∞. ˆ ˆ ˆ Therefore, f ∈ C0(R) and kfku = kfk∞ ≤ kfk1. Remarks. The Banach space C0(R) is a commutative Banach algebra when equipped with the pointwise product: kfgku ≤ kfkukgku. There is no unit in this C0(R), i.e. no element u ∈ C0(R) such that uf = f for all f ∈ C0(R). [The only candidate for a unit is the constant 1 function, which is continuous, but does not vanish at infinity]. Definition. Let A and B be (complex) Banach algebras. An algebra homomorphism is a linear transformation Λ : A → B such that Λ(fg) = Λ(f)Λ(g) for all f, g ∈ A. Notation. Let F denote the Fourier transform: F(f) = fˆ. 1 Proposition. F : L (R) → C0(R) is an algebra homomorphism for which (a) kFk ≤ 1 [the Fourier transform is a bounded linear transformation], and (b) kF(f ∗ g)ku ≤ kfk1kgk1 for all f, g. Proof. Linearity of F is clear, as is F(f ∗ g) = F(f)F(g). 1 By Theorem 9.6, kF(f)ku ≤ kfk1 for all f ∈ L (R), which implies that kFk = sup{kF(f)ku : kfk1 = 1} ≤ 1. This and the Convolution Theorem imply that kF(f ∗ g)ku ≤ kFk kf ∗ gk1 ≤ kf ∗ gk1 ≤ kfk1kgk1.//// Proposition. The commutative Banach algebra L1(R) has no unit. Proof. Suppose that there is a unit v ∈ L1(R), i.e. f ∗ v = f for all f ∈ L1(R). Since F is an algebra homomorphism, it follows that F(f) = F(f)F(v); hence fˆ(t) vˆ(t) − 1 = 0 for all t ∈ R. 1 Consider the L (R) function f(x) = χ[−1,1](x). For this, ( 2t−1 sin t if t 6= 0, fˆ(t) = F(f)(t) = 2 if t = 0. [see Homework Problem Ch.9 #2]. iαx ˆ ˆ ˆ ˆ Since F f(x)e = f(t − α) = fα(t) for α ∈ R, then fα(α) = f(0) = 2. ˆ Thus fα(α) vˆ(α) − 1 = 0 implies thatv ˆ(α) = 1 for each α ∈ R. But the constant 1 function is not in C0(R). .
Recommended publications
  • A Proof of Cantor's Theorem
    Cantor’s Theorem Joe Roussos 1 Preliminary ideas Two sets have the same number of elements (are equinumerous, or have the same cardinality) iff there is a bijection between the two sets. Mappings: A mapping, or function, is a rule that associates elements of one set with elements of another set. We write this f : X ! Y , f is called the function/mapping, the set X is called the domain, and Y is called the codomain. We specify what the rule is by writing f(x) = y or f : x 7! y. e.g. X = f1; 2; 3g;Y = f2; 4; 6g, the map f(x) = 2x associates each element x 2 X with the element in Y that is double it. A bijection is a mapping that is injective and surjective.1 • Injective (one-to-one): A function is injective if it takes each element of the do- main onto at most one element of the codomain. It never maps more than one element in the domain onto the same element in the codomain. Formally, if f is a function between set X and set Y , then f is injective iff 8a; b 2 X; f(a) = f(b) ! a = b • Surjective (onto): A function is surjective if it maps something onto every element of the codomain. It can map more than one thing onto the same element in the codomain, but it needs to hit everything in the codomain. Formally, if f is a function between set X and set Y , then f is surjective iff 8y 2 Y; 9x 2 X; f(x) = y Figure 1: Injective map.
    [Show full text]
  • Linear Transformation (Sections 1.8, 1.9) General View: Given an Input, the Transformation Produces an Output
    Linear Transformation (Sections 1.8, 1.9) General view: Given an input, the transformation produces an output. In this sense, a function is also a transformation. 1 4 3 1 3 Example. Let A = and x = 1 . Describe matrix-vector multiplication Ax 2 0 5 1 1 1 in the language of transformation. 1 4 3 1 31 5 Ax b 2 0 5 11 8 1 Vector x is transformed into vector b by left matrix multiplication Definition and terminologies. Transformation (or function or mapping) T from Rn to Rm is a rule that assigns to each vector x in Rn a vector T(x) in Rm. • Notation: T: Rn → Rm • Rn is the domain of T • Rm is the codomain of T • T(x) is the image of vector x • The set of all images T(x) is the range of T • When T(x) = Ax, A is a m×n size matrix. Range of T = Span{ column vectors of A} (HW1.8.7) See class notes for other examples. Linear Transformation --- Existence and Uniqueness Questions (Section 1.9) Definition 1: T: Rn → Rm is onto if each b in Rm is the image of at least one x in Rn. • i.e. codomain Rm = range of T • When solve T(x) = b for x (or Ax=b, A is the standard matrix), there exists at least one solution (Existence question). Definition 2: T: Rn → Rm is one-to-one if each b in Rm is the image of at most one x in Rn. • i.e. When solve T(x) = b for x (or Ax=b, A is the standard matrix), there exists either a unique solution or none at all (Uniqueness question).
    [Show full text]
  • Functions and Inverses
    Functions and Inverses CS 2800: Discrete Structures, Spring 2015 Sid Chaudhuri Recap: Relations and Functions ● A relation between sets A !the domain) and B !the codomain" is a set of ordered pairs (a, b) such that a ∈ A, b ∈ B !i.e. it is a subset o# A × B" Cartesian product – The relation maps each a to the corresponding b ● Neither all possible a%s, nor all possible b%s, need be covered – Can be one-one, one&'an(, man(&one, man(&man( Alice CS 2800 Bob A Carol CS 2110 B David CS 3110 Recap: Relations and Functions ● ) function is a relation that maps each element of A to a single element of B – Can be one-one or man(&one – )ll elements o# A must be covered, though not necessaril( all elements o# B – Subset o# B covered b( the #unction is its range/image Alice Balch Bob A Carol Jameson B David Mews Recap: Relations and Functions ● Instead of writing the #unction f as a set of pairs, e usually speci#y its domain and codomain as: f : A → B * and the mapping via a rule such as: f (Heads) = 0.5, f (Tails) = 0.5 or f : x ↦ x2 +he function f maps x to x2 Recap: Relations and Functions ● Instead of writing the #unction f as a set of pairs, e usually speci#y its domain and codomain as: f : A → B * and the mapping via a rule such as: f (Heads) = 0.5, f (Tails) = 0.5 2 or f : x ↦ x f(x) ● Note: the function is f, not f(x), – f(x) is the value assigned b( f the #unction f to input x x Recap: Injectivity ● ) function is injective (one-to-one) if every element in the domain has a unique i'age in the codomain – +hat is, f(x) = f(y) implies x = y Albany NY New York A MA Sacramento B CA Boston ..
    [Show full text]
  • Convolution.Pdf
    Convolution March 22, 2020 0.0.1 Setting up [23]: from bokeh.plotting import figure from bokeh.io import output_notebook, show,curdoc from bokeh.themes import Theme import numpy as np output_notebook() theme=Theme(json={}) curdoc().theme=theme 0.1 Convolution Convolution is a fundamental operation that appears in many different contexts in both pure and applied settings in mathematics. Polynomials: a simple case Let’s look at a simple case first. Consider the space of real valued polynomial functions P. If f 2 P, we have X1 i f(z) = aiz i=0 where ai = 0 for i sufficiently large. There is an obvious linear map a : P! c00(R) where c00(R) is the space of sequences that are eventually zero. i For each integer i, we have a projection map ai : c00(R) ! R so that ai(f) is the coefficient of z for f 2 P. Since P is a ring under addition and multiplication of functions, these operations must be reflected on the other side of the map a. 1 Clearly we can add sequences in c00(R) componentwise, and since addition of polynomials is also componentwise we have: a(f + g) = a(f) + a(g) What about multiplication? We know that X Xn X1 an(fg) = ar(f)as(g) = ar(f)an−r(g) = ar(f)an−r(g) r+s=n r=0 r=0 where we adopt the convention that ai(f) = 0 if i < 0. Given (a0; a1;:::) and (b0; b1;:::) in c00(R), define a ∗ b = (c0; c1;:::) where X1 cn = ajbn−j j=0 with the same convention that ai = bi = 0 if i < 0.
    [Show full text]
  • Tensor Products of F-Algebras
    Mediterr. J. Math. (2017) 14:63 DOI 10.1007/s00009-017-0841-x c The Author(s) 2017 This article is published with open access at Springerlink.com Tensor Products of f-algebras G. J. H. M. Buskes and A. W. Wickstead Abstract. We construct the tensor product for f-algebras, including proving a universal property for it, and investigate how it preserves algebraic properties of the factors. Mathematics Subject Classification. 06A70. Keywords. f-algebras, Tensor products. 1. Introduction An f-algebra is a vector lattice A with an associative multiplication , such that 1. a, b ∈ A+ ⇒ ab∈ A+ 2. If a, b, c ∈ A+ with a ∧ b = 0 then (ac) ∧ b =(ca) ∧ b =0. We are only concerned in this paper with Archimedean f-algebras so hence- forth, all f-algebras and, indeed, all vector lattices are assumed to be Archimedean. In this paper, we construct the tensor product of two f-algebras A and B. To be precise, we show that there is a natural way in which the Frem- lin Archimedean vector lattice tensor product A⊗B can be endowed with an f-algebra multiplication. This tensor product has the expected univer- sal property, at least as far as order bounded algebra homomorphisms are concerned. We also show that, unlike the case for order theoretic properties, algebraic properties behave well under this product. Again, let us emphasise here that algebra homomorphisms are not assumed to map an identity to an identity. Our approach is very much representational, however, much that might annoy some purists. In §2, we derive the representations needed, which are very mild improvements on results already in the literature.
    [Show full text]
  • Convolution Products of Monodiffric Functions
    View metadata, citation and similar papers at core.ac.uk brought to you by CORE provided by Elsevier - Publisher Connector JOC‘RNAL OF MATHEMATICAL ANALYSIS AND APPLICATIONS 37, 271-287 (1972) Convolution Products of Monodiffric Functions GEORGE BERZSENYI Lamar University, Beaumont, Texas 77710 Submitted by R. J. Dufin As it has been repeatedly pointed out by mathematicians working in discrete function theory, it is both essential and difficult to find discrete analogs for the ordinary pointwise product operation of analytic functions of a complex variable. R. P. Isaacs [Ill, the initiator of monodiffric function theory, is to be credited with the first results in this direction. Later G. J. Kurowski [13] introduced some new monodiffric analogues of pointwise multiplication while the present author [2] succeeded in sharpening and generalizing Isaacs’ and Kurowski’s results. However, all of these attempts were mainly concerned with trying to preserve as many of the pointwise aspects as possible. Consequently, each of these operations proved to be less than satisfactory either from the algebraic or from the function theoretic viewpoint. In the present paper, a different approach is taken. Influenced by the suc- cess of R. J. Duffin and C. S. Duris [5] in a related area, we utilize the line integrals introduced in Ref. [l] and the conjoint line integral of Isaacs [l I] to define several convolution-type product operations. It is shown that if D is a suitably restricted discrete domain then monodiffricity is preserved by each of these operations. Further algebraic and function theoretic properties of these operations are investigated and some examples of their applicability are given.
    [Show full text]
  • Math 101 B-Packet
    Math 101 B-Packet Scott Rome Winter 2012-13 1 Redefining functions This quarter we have defined a function as a rule which assigns exactly one output to each input, and so far we have been happy with this definition. Unfortunately, this way of thinking of a function is insufficient as things become more complicated in mathematics. For a better understanding of a function, we will first need to define it better. Definition 1.1. Let X; Y be any sets. A function f : X ! Y is a rule which assigns every element of X to an element of Y . The sets X and Y are called the domain and codomain of f respectively. x f(x) y Figure 1: This function f : X ! Y maps x 7! f(x). The green circle indicates the range of the function. Notice y is in the codomain, but f does not map to it. Remark 1.2. It is necessary to define the rule, the domain, and the codomain to define a function. Thus far in the class, we have been \sloppy" when working with functions. Remark 1.3. Notice how in the definition, the function is defined by three things: the rule, the domain, and the codomain. That means you can define functions that seem to be the same, but are actually different as we will see. The domain of a function can be thought of as the set of all inputs (that is, everything in the domain will be mapped somewhere by the function). On the other hand, the codomain of a function is the set of all possible outputs, and a function may not necessarily map to every element of the codomain.
    [Show full text]
  • Sets, Functions, and Domains
    Sets, Functions, and Domains Functions are fundamental to denotational semantics. This chapter introduces functions through set theory, which provides a precise yet intuitive formulation. In addition, the con- cepts of set theory form a foundation for the theory of semantic domains, the value spaces used for giving meaning to languages. We examine the basic principles of sets, functions, and domains in turn. 2.1 SETS __________________________________________________________________________________________________________________________________ A set is a collection; it can contain numbers, persons, other sets, or (almost) anything one wishes. Most of the examples in this book use numbers and sets of numbers as the members of sets. Like any concept, a set needs a representation so that it can be written down. Braces are used to enclose the members of a set. Thus, { 1, 4, 7 } represents the set containing the numbers 1, 4, and 7. These are also sets: { 1, { 1, 4, 7 }, 4 } { red, yellow, grey } {} The last example is the empty set, the set with no members, also written as ∅. When a set has a large number of members, it is more convenient to specify the condi- tions for membership than to write all the members. A set S can be de®ned by S= { x | P(x) }, which says that an object a belongs to S iff (if and only if) a has property P, that is, P(a) holds true. For example, let P be the property ``is an even integer.'' Then { x | x is an even integer } de®nes the set of even integers, an in®nite set. Note that ∅ can be de®ned as the set { x | x≠x }.
    [Show full text]
  • Parametrised Enumeration
    Parametrised Enumeration Arne Meier Februar 2020 color guide (the dark blue is the same with the Gottfried Wilhelm Leibnizuniversity Universität logo) Hannover c: 100 m: 70 Institut für y: 0 k: 0 Theoretische c: 35 m: 0 y: 10 Informatik k: 0 c: 0 m: 0 y: 0 k: 35 Fakultät für Elektrotechnik und Informatik Institut für Theoretische Informatik Fachgebiet TheoretischeWednesday, February 18, 15 Informatik Habilitationsschrift Parametrised Enumeration Arne Meier geboren am 6. Mai 1982 in Hannover Februar 2020 Gutachter Till Tantau Institut für Theoretische Informatik Universität zu Lübeck Gutachter Stefan Woltran Institut für Logic and Computation 192-02 Technische Universität Wien Arne Meier Parametrised Enumeration Habilitationsschrift, Datum der Annahme: 20.11.2019 Gutachter: Till Tantau, Stefan Woltran Gottfried Wilhelm Leibniz Universität Hannover Fachgebiet Theoretische Informatik Institut für Theoretische Informatik Fakultät für Elektrotechnik und Informatik Appelstrasse 4 30167 Hannover Dieses Werk ist lizenziert unter einer Creative Commons “Namensnennung-Nicht kommerziell 3.0 Deutschland” Lizenz. Für Julia, Jonas Heinrich und Leonie Anna. Ihr seid mein größtes Glück auf Erden. Danke für eure Geduld, euer Verständnis und eure Unterstützung. Euer Rückhalt bedeutet mir sehr viel. ♥ If I had an hour to solve a problem, I’d spend 55 minutes thinking about the problem and 5 min- “ utes thinking about solutions. ” — Albert Einstein Abstract In this thesis, we develop a framework of parametrised enumeration complexity. At first, we provide the reader with preliminary notions such as machine models and complexity classes besides proving them to be well-chosen. Then, we study the interplay and the landscape of these classes and present connections to classical enumeration classes.
    [Show full text]
  • Vector Space Theory
    Vector Space Theory A course for second year students by Robert Howlett typesetting by TEX Contents Copyright University of Sydney Chapter 1: Preliminaries 1 §1a Logic and commonSchool sense of Mathematics and Statistics 1 §1b Sets and functions 3 §1c Relations 7 §1d Fields 10 Chapter 2: Matrices, row vectors and column vectors 18 §2a Matrix operations 18 §2b Simultaneous equations 24 §2c Partial pivoting 29 §2d Elementary matrices 32 §2e Determinants 35 §2f Introduction to eigenvalues 38 Chapter 3: Introduction to vector spaces 49 §3a Linearity 49 §3b Vector axioms 52 §3c Trivial consequences of the axioms 61 §3d Subspaces 63 §3e Linear combinations 71 Chapter 4: The structure of abstract vector spaces 81 §4a Preliminary lemmas 81 §4b Basis theorems 85 §4c The Replacement Lemma 86 §4d Two properties of linear transformations 91 §4e Coordinates relative to a basis 93 Chapter 5: Inner Product Spaces 99 §5a The inner product axioms 99 §5b Orthogonal projection 106 §5c Orthogonal and unitary transformations 116 §5d Quadratic forms 121 iii Chapter 6: Relationships between spaces 129 §6a Isomorphism 129 §6b Direct sums Copyright University of Sydney 134 §6c Quotient spaces 139 §6d The dual spaceSchool of Mathematics and Statistics 142 Chapter 7: Matrices and Linear Transformations 148 §7a The matrix of a linear transformation 148 §7b Multiplication of transformations and matrices 153 §7c The Main Theorem on Linear Transformations 157 §7d Rank and nullity of matrices 161 Chapter 8: Permutations and determinants 171 §8a Permutations 171 §8b Determinants
    [Show full text]
  • Categories and Computability
    Categories and Computability J.R.B. Cockett ∗ Department of Computer Science, University of Calgary, Calgary, T2N 1N4, Alberta, Canada February 22, 2010 Contents 1 Introduction 3 1.1 Categories and foundations . ........ 3 1.2 Categoriesandcomputerscience . ......... 4 1.3 Categoriesandcomputability . ......... 4 I Basic Categories 6 2 Categories and examples 7 2.1 Thedefinitionofacategory . ....... 7 2.1.1 Categories as graphs with composition . ......... 7 2.1.2 Categories as partial semigroups . ........ 8 2.1.3 Categoriesasenrichedcategories . ......... 9 2.1.4 The opposite category and duality . ....... 9 2.2 Examplesofcategories. ....... 10 2.2.1 Preorders ..................................... 10 2.2.2 Finitecategories .............................. 11 2.2.3 Categoriesenrichedinfinitesets . ........ 11 2.2.4 Monoids....................................... 13 2.2.5 Pathcategories................................ 13 2.2.6 Matricesoverarig .............................. 13 2.2.7 Kleenecategories.............................. 14 2.2.8 Sets ......................................... 15 2.2.9 Programminglanguages . 15 2.2.10 Productsandsumsofcategories . ....... 16 2.2.11 Slicecategories .............................. ..... 16 ∗Partially supported by NSERC, Canada. 1 3 Elements of category theory 18 3.1 Epics, monics, retractions, and sections . ............. 18 3.2 Functors........................................ 20 3.3 Natural transformations . ........ 23 3.4 Adjoints........................................ 25 3.4.1 Theuniversalproperty.
    [Show full text]
  • The Entscheidungsproblem and Alan Turing
    The Entscheidungsproblem and Alan Turing Author: Laurel Brodkorb Advisor: Dr. Rachel Epstein Georgia College and State University December 18, 2019 1 Abstract Computability Theory is a branch of mathematics that was developed by Alonzo Church, Kurt G¨odel,and Alan Turing during the 1930s. This paper explores their work to formally define what it means for something to be computable. Most importantly, this paper gives an in-depth look at Turing's 1936 paper, \On Computable Numbers, with an Application to the Entscheidungsproblem." It further explores Turing's life and impact. 2 Historic Background Before 1930, not much was defined about computability theory because it was an unexplored field (Soare 4). From the late 1930s to the early 1940s, mathematicians worked to develop formal definitions of computable functions and sets, and applying those definitions to solve logic problems, but these problems were not new. It was not until the 1800s that mathematicians began to set an axiomatic system of math. This created problems like how to define computability. In the 1800s, Cantor defined what we call today \naive set theory" which was inconsistent. During the height of his mathematical period, Hilbert defended Cantor but suggested an axiomatic system be established, and characterized the Entscheidungsproblem as the \fundamental problem of mathemat- ical logic" (Soare 227). The Entscheidungsproblem was proposed by David Hilbert and Wilhelm Ackerman in 1928. The Entscheidungsproblem, or Decision Problem, states that given all the axioms of math, there is an algorithm that can tell if a proposition is provable. During the 1930s, Alonzo Church, Stephen Kleene, Kurt G¨odel,and Alan Turing worked to formalize how we compute anything from real numbers to sets of functions to solve the Entscheidungsproblem.
    [Show full text]