The Dirac Delta Function

Total Page:16

File Type:pdf, Size:1020Kb

The Dirac Delta Function Appendix A The Dirac Delta Function A.1 The One-Dimensional Dirac Delta Function The Dirac delta function [1] in one-dimensional space may be defined by the pair of equations δ(x) = 0; x = 0, (A.1) ∞ δ(x)dx = 1. (A.2) −∞ It is clear from this definition that δ(x) is not a function in the ordinary mathematical sense, because if a function is zero everywhere except at a single point and the integral of this function over its entire domain of definition exists, then the value of this integral is necessarily also equal to zero. Because of this, it is more appropriate to regard δ(x) as a functional quantity with a certain well-defined symbolic meaning. For example, one can consider a sequence of functions δ(x,ε) that, with increasing values of the parameter ε, differ appreciably from zero only over a decreasing x-interval about the origin and which are such that ∞ δ(x,ε)dx = 1(A.3) −∞ for all values of ε. Although it may be tempting to try to interpret the Dirac delta function as the limit of such a sequence of well-defined functions δ(x,ε) as ε →∞, it must be recognized that this limit need not exist for all values of the independent variable x. However, the limit ∞ lim δ(x,ε)dx = 1(A.4) ε→∞ −∞ © Springer Nature Switzerland AG 2019 641 K. E. Oughstun, Electromagnetic and Optical Pulse Propagation, Springer Series in Optical Sciences 224, https://doi.org/10.1007/978-3-030-20835-6 642 A The Dirac Delta Function must exist. As a consequence, one may interpret any operation that involves the delta function δ(x) as implying that this operation is to be performed with a function δ(x,ε) of a suitable sequence and that the limit as ε →∞is to be taken at the conclusion of the calculation. The particular choice of the sequence of functions δ(x,ε) is immaterial, provided that any oscillations near the origin x = 0 are not too violent [2]. For example, each of the following functions forms a sequence with respect to the parameter ε that satisfies the required properties: ε −ε2x2 ε δ(x,ε) = √ e & δ(x,ε) = rect /ε(x) & δ(x,ε) = sinc(εx) π 1 π in Cartesian coordinates, and ε δ(r,ε) = circ (r) & δ(r,ε) = J (2πεr) 1/ε r 1 in polar coordinates, where rect1/ε(x) ≡ ε/2 when |x| < 1/ε and is zero otherwise, 2 circ1/ε(r) ≡ ε /π when r<1/ε and is zero otherwise, and sinc(x) ≡ sin (x)/x when x = 0 and sinc(0) ≡ limx→0 sin (x)/x = 1. Let f(x) be a continuous and sufficiently well-behaved function of x ∈ (−∞, ∞) and consider the value of the definite integral ∞ ∞ f(x)δ(x− a)dx = lim f(x)δ(x− a,ε)dx. −∞ ε→∞ −∞ When the parameter ε is large, the value of the integral appearing on the right- hand side of this equation depends essentially on the behavior of f(x) in the immediate neighborhood of the point x = a alone, and the error that results from the replacement of f(x) by f(a) may be made as small as desired by taking ε sufficiently large. Hence ∞ ∞ lim f(x)δ(x− a,ε)dx = f(a) lim δ(x − a,ε)dx, ε→∞ −∞ ε→∞ −∞ so that ∞ f(x)δ(x− a)dx = f(a). (A.5) −∞ This result is referred to as the sifting property of the delta function. Notice that, for this result to hold, the domain of integration need not be extended over all x ∈ (−∞, ∞); it is only necessary that the domain of integration contain the point x = a A.1 The One-Dimensional Dirac Delta Function 643 in its interior, so that a+Δ2 f(x)δ(x− a)dx = f(a), (A.6) a−Δ1 where Δ1,2 > 0. It is then seen that f(x)need only be continuous at x = a. The above results may be written symbolically as f(x)δ(x− a) = f(a)δ(x− a), (A.7) the meaning of such a statement being that the two sides yield the same result when integrated over any domain containing the point x = a. For the special case when f(x)= xk with k>0 and a = 0, Eq. (A.7) yields xkδ(x) = 0, ∀k>0. (A.8) Theorem A.1 (Similarity Relationship (Scaling Law)) For all a = 0 1 δ(ax) = δ(x). (A.9) |a| Proof In order to prove this relationship one need only compare the integrals of f(x)δ(ax)and f(x)δ(x)/|a| for any sufficiently well-behaved continuous function f(x). For the first integral one has (for any a = 0) ∞ 1 ∞ 1 f(x)δ(ax)dx =± f(y/a)δ(y)dy = f(0), −∞ a −∞ |a| where the upper or lower sign choice is taken accordingly as a>0ora<0, respectively, and for the second integral one obtains ∞ 1 1 f(x) δ(x)dx = f(0). −∞ |a| |a| Comparison of these two results then shows that δ(ax) = δ(x)/|a|, as was to be proved. For the special case a =−1, Eq. (A.9) yields δ(−x) = δ(x), (A.10) so that the delta function is an even function of its argument. 644 A The Dirac Delta Function Theorem A.2 (Composite Function Theorem) If y = f(x) is any continuous function of x with simple zeroes at the points xi [i.e., y = 0 at x = xi and f (xi) = 0] and no other zeroes, then 1 δ(f(x)) = δ(x − x ). (A.11) |f (x )| i i i Proof Let g(x) be any sufficiently well-behaved continuous function and let {xi} denote the set of points at which y = 0. Under the change of variable x = f −1(y) one has that ∞ − 1 g(x)δ(f(x))dx = g f 1(y) δ(y) dy −1 −∞ R |f f (y) | − 1 1 = g f 1(0) = g(x ) , −1 i |f f (0) | |f (xi| xi i where R denotes the range of f(x). In addition, ∞ ∞ 1 − = 1 − g(x) δ(x xi) dx g(x)δ(x xi)dx −∞ |f (x )| |f (x )| −∞ i i i i 1 = g(x ) . i |f (x | i i Comparison of these two expressions then proves the theorem. As an example, consider the function f(x) = x2 − a2 which has simple zeroes at x =±a. Then |f (±a)|=2|a| so that, for a = 0, 1 δ(x2 − a2) = (δ(x − a) + δ(x + a)) . 2|a| An additional relationship of interest that employs the Dirac delta function is ∞ δ(ξ − x)δ(x − η)dx = δ(ξ − η), (A.12) −∞ which is seen to be an extension of the sifting property to the delta function itself. This equation then implies that if both sides are multiplied by a continuous function of either ξ or η and the result integrated over all values of either ξ or η, respectively, an identity is obtained. That is, because ∞ f(ξ)δ(ξ − η)dξ = f(η), −∞ A.1 The One-Dimensional Dirac Delta Function 645 and ∞ ∞ f(ξ) δ(ξ − x)δ(x − η)dx dξ −∞ −∞ ∞ ∞ = f(ξ)δ(ξ − x)dξ δ(x − η)dx −∞ −∞ ∞ = f(x)δ(x− η)dx = f(η) −∞ then the expression in Eq. (A.12) follows. In a similar manner, because ∞ ∞ g(η) δ(ξ − x)δ(x − η)dx dη −∞ −∞ ∞ ∞ = g(η)δ(x − η)dη δ(ξ − x)dx −∞ −∞ ∞ = g(x)δ(ξ − x)dx = g(ξ) −∞ and ∞ g(η)δ(ξ − η)dη = g(ξ), −∞ then the expression in Eq. (A.12) is again obtained. Consider next what interpretation may be given to the derivatives of the delta function, accomplished through use of the function sequence δ(x,ε). Consider then ∞ the ordinary integral −∞ f(x)δ(x, ε)dx which may be evaluated by application of the method of integration by parts with u = f(x)and dv = δ (x, ε)dx, so that ∞ ∞ f(x)δ(x, ε)dx = f(∞)δ(∞,ε)− f(−∞)δ(−∞,ε)− f (x)δ(x, ε)dx. −∞ −∞ Upon proceeding to the limit as ε →∞, the first two terms appearing on the right- hand side of this equation both vanish because lim δ(±∞,ε)= 0, (A.13) ε→∞ with the result ∞ f(x)δ(x)dx =−f (0). (A.14) −∞ 646 A The Dirac Delta Function Upon repeating this procedure n times for the nth-order derivative of the delta function, one obtains the general result ∞ f(x)δ(n)(x)dx = (−1)nf (n)(0). (A.15) −∞ As a special case of Eq. (A.14), let f(x)= x so that ∞ ∞ xδ (x)dx =−1 =− δ(x)dx, −∞ −∞ and one then has the equivalence xδ (x) =−δ(x). (A.16) Because δ(x) is an even function and x is an odd function, it then follows that δ (x) is an odd function of its argument; that is δ (−x) =−δ (x). (A.17) The generalization of Eq. (A.16) may be directly obtained from Eq. (A.15) by letting f(x)= xn. In that case, f (n)(x) = n! and this relation gives ∞ ∞ xnδ (x)dx = (−1)nn!=(−1)nn! δ(x)dx, −∞ −∞ and one then has the general equivalence xnδ(n)(x) = (−1)nn!δ(x). (A.18) The even-order derivatives of the delta function are even functions and the odd-order derivatives are odd functions of the argument.
Recommended publications
  • Distributions:The Evolutionof a Mathematicaltheory
    Historia Mathematics 10 (1983) 149-183 DISTRIBUTIONS:THE EVOLUTIONOF A MATHEMATICALTHEORY BY JOHN SYNOWIEC INDIANA UNIVERSITY NORTHWEST, GARY, INDIANA 46408 SUMMARIES The theory of distributions, or generalized func- tions, evolved from various concepts of generalized solutions of partial differential equations and gener- alized differentiation. Some of the principal steps in this evolution are described in this paper. La thgorie des distributions, ou des fonctions g&&alis~es, s'est d&eloppeg 2 partir de divers con- cepts de solutions g&&alis6es d'gquations aux d&i- &es partielles et de diffgrentiation g&&alis6e. Quelques-unes des principales &apes de cette &olution sont d&rites dans cet article. Die Theorie der Distributionen oder verallgemein- erten Funktionen entwickelte sich aus verschiedenen Begriffen von verallgemeinerten Lasungen der partiellen Differentialgleichungen und der verallgemeinerten Ableitungen. In diesem Artikel werden einige der wesentlichen Schritte dieser Entwicklung beschrieben. 1. INTRODUCTION The founder of the theory of distributions is rightly con- sidered to be Laurent Schwartz. Not only did he write the first systematic paper on the subject [Schwartz 19451, which, along with another paper [1947-19481, already contained most of the basic ideas, but he also wrote expository papers on distributions for electrical engineers [19481. Furthermore, he lectured effec- tively on his work, for example, at the Canadian Mathematical Congress [1949]. He showed how to apply distributions to partial differential equations [19SObl and wrote the first general trea- tise on the subject [19SOa, 19511. Recognition of his contri- butions came in 1950 when he was awarded the Fields' Medal at the International Congress of Mathematicians, and in his accep- tance address to the congress Schwartz took the opportunity to announce his celebrated "kernel theorem" [195Oc].
    [Show full text]
  • Chapter 3 Constitutive Relations
    Chapter 3 Constitutive Relations Maxwell’s equations define the fields that are generated by currents and charges. However, they do not describe how these currents and charges are generated. Thus, to find a self-consistent solution for the electromagnetic field, Maxwell’s equations must be supplemented by relations that describe the behavior of matter under the influence of fields. These material equations are known as constitutive relations. The constitutive relations express the secondary sources P and M in terms of the fields E and H, that is P = f[E] and M = f[H].1 According to Eq. (1.19) this is equivalent to D = f[E] and B = f[H]. If we expand these relations into power series ∂D 1 ∂2D D = D + E + E2 + .. (3.1) 0 ∂E 2 ∂E2 E=0 E=0 we find that the lowest-order term depends linearly on E. In most practical situ- ations the interaction of radiation with matter is weak and it suffices to truncate the power series after the linear term. The nonlinear terms come into play when the fields acting on matter become comparable to the atomic Coulomb potential. This is the territory of strong field physics. Here we will entirely focus on the linear properties of matter. 1In some exotic cases we can have P = f[E, H] and M = f[H, E], which are so-called bi- isotropic or bi-anisotropic materials. These are special cases and won’t be discussed here. 33 34 CHAPTER 3. CONSTITUTIVE RELATIONS 3.1 Linear Materials The most general linear relationship between D and E can be written as2 ∞ ∞ D(r, t) = ε ε˜(r r′, t t′) E(r′, t′) d3r′ dt′ , (3.2) 0 − − −∞Z−∞Z which states that the response D at the location r and at time t not only depends on the excitation E at r and t, but also on the excitation E in all other locations r′ and all other times t′.
    [Show full text]
  • The Green-Function Transform and Wave Propagation
    1 The Green-function transform and wave propagation Colin J. R. Sheppard,1 S. S. Kou2 and J. Lin2 1Department of Nanophysics, Istituto Italiano di Tecnologia, Genova 16163, Italy 2School of Physics, University of Melbourne, Victoria 3010, Australia *Corresponding author: [email protected] PACS: 0230Nw, 4120Jb, 4225Bs, 4225Fx, 4230Kq, 4320Bi Abstract: Fourier methods well known in signal processing are applied to three- dimensional wave propagation problems. The Fourier transform of the Green function, when written explicitly in terms of a real-valued spatial frequency, consists of homogeneous and inhomogeneous components. Both parts are necessary to result in a pure out-going wave that satisfies causality. The homogeneous component consists only of propagating waves, but the inhomogeneous component contains both evanescent and propagating terms. Thus we make a distinction between inhomogenous waves and evanescent waves. The evanescent component is completely contained in the region of the inhomogeneous component outside the k-space sphere. Further, propagating waves in the Weyl expansion contain both homogeneous and inhomogeneous components. The connection between the Whittaker and Weyl expansions is discussed. A list of relevant spherically symmetric Fourier transforms is given. 2 1. Introduction In an impressive recent paper, Schmalz et al. presented a rigorous derivation of the general Green function of the Helmholtz equation based on three-dimensional (3D) Fourier transformation, and then found a unique solution for the case of a source (Schmalz, Schmalz et al. 2010). Their approach is based on the use of generalized functions and the causal nature of the out-going Green function. Actually, the basic principle of their method was described many years ago by Dirac (Dirac 1981), but has not been widely adopted.
    [Show full text]
  • Approximate Separability of Green's Function for High Frequency
    APPROXIMATE SEPARABILITY OF GREEN'S FUNCTION FOR HIGH FREQUENCY HELMHOLTZ EQUATIONS BJORN¨ ENGQUIST AND HONGKAI ZHAO Abstract. Approximate separable representations of Green's functions for differential operators is a basic and an important aspect in the analysis of differential equations and in the development of efficient numerical algorithms for solving them. Being able to approx- imate a Green's function as a sum with few separable terms is equivalent to the existence of low rank approximation of corresponding discretized system. This property can be explored for matrix compression and efficient numerical algorithms. Green's functions for coercive elliptic differential operators in divergence form have been shown to be highly separable and low rank approximation for their discretized systems has been utilized to develop efficient numerical algorithms. The case of Helmholtz equation in the high fre- quency limit is more challenging both mathematically and numerically. In this work, we develop a new approach to study approximate separability for the Green's function of Helmholtz equation in the high frequency limit based on an explicit characterization of the relation between two Green's functions and a tight dimension estimate for the best linear subspace approximating a set of almost orthogonal vectors. We derive both lower bounds and upper bounds and show their sharpness for cases that are commonly used in practice. 1. Introduction Given a linear differential operator, denoted by L, the Green's function, denoted by G(x; y), is defined as the fundamental solution in a domain Ω ⊆ Rn to the partial differ- ential equation 8 n < LxG(x; y) = δ(x − y); x; y 2 Ω ⊆ R (1) : with boundary condition or condition at infinity; where δ(x − y) is the Dirac delta function denoting an impulse source point at y.
    [Show full text]
  • Arxiv:1404.7630V2 [Math.AT] 5 Nov 2014 Asoffsae,Se[S4 P8,Vr6.Isedof Instead Ver66]
    PROPER BASE CHANGE FOR SEPARATED LOCALLY PROPER MAPS OLAF M. SCHNÜRER AND WOLFGANG SOERGEL Abstract. We introduce and study the notion of a locally proper map between topological spaces. We show that fundamental con- structions of sheaf theory, more precisely proper base change, pro- jection formula, and Verdier duality, can be extended from contin- uous maps between locally compact Hausdorff spaces to separated locally proper maps between arbitrary topological spaces. Contents 1. Introduction 1 2. Locally proper maps 3 3. Proper direct image 5 4. Proper base change 7 5. Derived proper direct image 9 6. Projection formula 11 7. Verdier duality 12 8. The case of unbounded derived categories 15 9. Remindersfromtopology 17 10. Reminders from sheaf theory 19 11. Representability and adjoints 21 12. Remarks on derived functors 22 References 23 arXiv:1404.7630v2 [math.AT] 5 Nov 2014 1. Introduction The proper direct image functor f! and its right derived functor Rf! are defined for any continuous map f : Y → X of locally compact Hausdorff spaces, see [KS94, Spa88, Ver66]. Instead of f! and Rf! we will use the notation f(!) and f! for these functors. They are embedded into a whole collection of formulas known as the six-functor-formalism of Grothendieck. Under the assumption that the proper direct image functor f(!) has finite cohomological dimension, Verdier proved that its ! derived functor f! admits a right adjoint f . Olaf Schnürer was supported by the SPP 1388 and the SFB/TR 45 of the DFG. Wolfgang Soergel was supported by the SPP 1388 of the DFG. 1 2 OLAFM.SCHNÜRERANDWOLFGANGSOERGEL In this article we introduce in 2.3 the notion of a locally proper map between topological spaces and show that the above results even hold for arbitrary topological spaces if all maps whose proper direct image functors are involved are locally proper and separated.
    [Show full text]
  • B1. Fourier Analysis of Discrete Time Signals
    B1. Fourier Analysis of Discrete Time Signals Objectives • Introduce discrete time periodic signals • Define the Discrete Fourier Series (DFS) expansion of periodic signals • Define the Discrete Fourier Transform (DFT) of signals with finite length • Determine the Discrete Fourier Transform of a complex exponential 1. Introduction In the previous chapter we defined the concept of a signal both in continuous time (analog) and discrete time (digital). Although the time domain is the most natural, since everything (including our own lives) evolves in time, it is not the only possible representation. In this chapter we introduce the concept of expanding a generic signal in terms of elementary signals, such as complex exponentials and sinusoids. This leads to the frequency domain representation of a signal in terms of its Fourier Transform and the concept of frequency spectrum so that we characterize a signal in terms of its frequency components. First we begin with the introduction of periodic signals, which keep repeating in time. For these signals it is fairly easy to determine an expansion in terms of sinusoids and complex exponentials, since these are just particular cases of periodic signals. This is extended to signals of a finite duration which becomes the Discrete Fourier Transform (DFT), one of the most widely used algorithms in Signal Processing. The concepts introduced in this chapter are at the basis of spectral estimation of signals addressed in the next two chapters. 2. Periodic Signals VIDEO: Periodic Signals (19:45) http://faculty.nps.edu/rcristi/eo3404/b-discrete-fourier-transform/videos/chapter1-seg1_media/chapter1-seg1- 0.wmv http://faculty.nps.edu/rcristi/eo3404/b-discrete-fourier-transform/videos/b1_02_periodicSignals.mp4 In this section we define a class of discrete time signals called Periodic Signals.
    [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]
  • 7: Inner Products, Fourier Series, Convolution
    7: FOURIER SERIES STEVEN HEILMAN Contents 1. Review 1 2. Introduction 1 3. Periodic Functions 2 4. Inner Products on Periodic Functions 3 5. Trigonometric Polynomials 5 6. Periodic Convolutions 7 7. Fourier Inversion and Plancherel Theorems 10 8. Appendix: Notation 13 1. Review Exercise 1.1. Let (X; h·; ·i) be a (real or complex) inner product space. Define k·k : X ! [0; 1) by kxk := phx; xi. Then (X; k·k) is a normed linear space. Consequently, if we define d: X × X ! [0; 1) by d(x; y) := ph(x − y); (x − y)i, then (X; d) is a metric space. Exercise 1.2 (Cauchy-Schwarz Inequality). Let (X; h·; ·i) be a complex inner product space. Let x; y 2 X. Then jhx; yij ≤ hx; xi1=2hy; yi1=2: Exercise 1.3. Let (X; h·; ·i) be a complex inner product space. Let x; y 2 X. As usual, let kxk := phx; xi. Prove Pythagoras's theorem: if hx; yi = 0, then kx + yk2 = kxk2 + kyk2. 1 Theorem 1.4 (Weierstrass M-test). Let (X; d) be a metric space and let (fj)j=1 be a sequence of bounded real-valued continuous functions on X such that the series (of real num- P1 P1 bers) j=1 kfjk1 is absolutely convergent. Then the series j=1 fj converges uniformly to some continuous function f : X ! R. 2. Introduction A general problem in analysis is to approximate a general function by a series that is relatively easy to describe. With the Weierstrass Approximation theorem, we saw that it is possible to achieve this goal by approximating compactly supported continuous functions by polynomials.
    [Show full text]
  • Linear Differential Equations in N Th-Order ;
    Ordinary Differential Equations Second Course Babylon University 2016-2017 Linear Differential Equations in n th-order ; An nth-order linear differential equation has the form ; …. (1) Where the coefficients ;(j=0,1,2,…,n-1) ,and depends solely on the variable (not on or any derivative of ) If then Eq.(1) is homogenous ,if not then is nonhomogeneous . …… 2 A linear differential equation has constant coefficients if all the coefficients are constants ,if one or more is not constant then has variable coefficients . Examples on linear differential equations ; first order nonhomogeneous second order homogeneous third order nonhomogeneous fifth order homogeneous second order nonhomogeneous Theorem 1: Consider the initial value problem given by the linear differential equation(1)and the n initial Conditions; … …(3) Define the differential operator L(y) by …. (4) Where ;(j=0,1,2,…,n-1) is continuous on some interval of interest then L(y) = and a linear homogeneous differential equation written as; L(y) =0 Definition: Linearly Independent Solution; A set of functions … is linearly dependent on if there exists constants … not all zero ,such that …. (5) on 1 Ordinary Differential Equations Second Course Babylon University 2016-2017 A set of functions is linearly dependent if there exists another set of constants not all zero, that is satisfy eq.(5). If the only solution to eq.(5) is ,then the set of functions is linearly independent on The functions are linearly independent Theorem 2: If the functions y1 ,y2, …, yn are solutions for differential
    [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]
  • Fourier Series
    Academic Press Encyclopedia of Physical Science and Technology Fourier Series James S. Walker Department of Mathematics University of Wisconsin–Eau Claire Eau Claire, WI 54702–4004 Phone: 715–836–3301 Fax: 715–836–2924 e-mail: [email protected] 1 2 Encyclopedia of Physical Science and Technology I. Introduction II. Historical background III. Definition of Fourier series IV. Convergence of Fourier series V. Convergence in norm VI. Summability of Fourier series VII. Generalized Fourier series VIII. Discrete Fourier series IX. Conclusion GLOSSARY ¢¤£¦¥¨§ Bounded variation: A function has bounded variation on a closed interval ¡ ¢ if there exists a positive constant © such that, for all finite sets of points "! "! $#&% (' #*) © ¥ , the inequality is satisfied. Jordan proved that a function has bounded variation if and only if it can be expressed as the difference of two non-decreasing functions. Countably infinite set: A set is countably infinite if it can be put into one-to-one £0/"£ correspondence with the set of natural numbers ( +,£¦-.£ ). Examples: The integers and the rational numbers are countably infinite sets. "! "!;: # # 123547698 Continuous function: If , then the function is continuous at the point : . Such a point is called a continuity point for . A function which is continuous at all points is simply referred to as continuous. Lebesgue measure zero: A set < of real numbers is said to have Lebesgue measure ! $#¨CED B ¢ £¦¥ zero if, for each =?>A@ , there exists a collection of open intervals such ! ! D D J# K% $#L) ¢ £¦¥ ¥ ¢ = that <GFIH and . Examples: All finite sets, and all countably infinite sets, have Lebesgue measure zero. "! "! % # % # Odd and even functions: A function is odd if for all in its "! "! % # # domain.
    [Show full text]
  • Introduction Into Quaternions for Spacecraft Attitude Representation
    Introduction into quaternions for spacecraft attitude representation Dipl. -Ing. Karsten Groÿekatthöfer, Dr. -Ing. Zizung Yoon Technical University of Berlin Department of Astronautics and Aeronautics Berlin, Germany May 31, 2012 Abstract The purpose of this paper is to provide a straight-forward and practical introduction to quaternion operation and calculation for rigid-body attitude representation. Therefore the basic quaternion denition as well as transformation rules and conversion rules to or from other attitude representation parameters are summarized. The quaternion computation rules are supported by practical examples to make each step comprehensible. 1 Introduction Quaternions are widely used as attitude represenation parameter of rigid bodies such as space- crafts. This is due to the fact that quaternion inherently come along with some advantages such as no singularity and computationally less intense compared to other attitude parameters such as Euler angles or a direction cosine matrix. Mainly, quaternions are used to • Parameterize a spacecraft's attitude with respect to reference coordinate system, • Propagate the attitude from one moment to the next by integrating the spacecraft equa- tions of motion, • Perform a coordinate transformation: e.g. calculate a vector in body xed frame from a (by measurement) known vector in inertial frame. However, dierent references use several notations and rules to represent and handle attitude in terms of quaternions, which might be confusing for newcomers [5], [4]. Therefore this article gives a straight-forward and clearly notated introduction into the subject of quaternions for attitude representation. The attitude of a spacecraft is its rotational orientation in space relative to a dened reference coordinate system.
    [Show full text]