fourier series - georgi p.tolstov (dover 1976, 349s)

349

Click here to load reader

Upload: alancardekpa

Post on 26-Oct-2014

1.509 views

Category:

Documents


80 download

TRANSCRIPT

Page 1: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)
Page 2: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

DOVER BOOKS ON MATHEMATICS ARITHMETIC REFRESHER, A. Albert Klaf. (2 124 1-6) $8.95 CALCULUS REFRESHER, A. A1 bert Klaf. (20370-0) $8.95 THEORY OF APPROXIMATION, N.1. Achieser. (67129-1) $8.95 ON FORMALLY UNDECIDABLE PROPOSITIONS OF PRINCIPIA MATHEMAT~CA AND

RELATED SYSTEMS, Kurt Godel. (66980-7) $5.95 METHODS OF APPLIED MATHEMATICS, Francis B. Hildebrand. (67002-3) $10.95 MATHEMATICS AND LOGIC, Mark Kac & Stanislaw M. Ulam. (67085-6) $7.95 MATHEMATICAL METHODS AND THEORY IN GAMES, PROGRAMMING, AND ECONOMICS,

Samuel Karlin. (67020-1) $18.95 SURVEY OF MATRIX THEORY AND MATRIX INEQUALITIES, Marvin Marcus &

Henryk Minc. (67102-X) $6.95 A~STRACT ALGEBRA AND SOLUTION BY RADICALS,. John E. and Margaret W.

Maxfield. (67121-6) $8.95 ELEMENTS OF THE TOPOLOGY OF PLANE SETS OF POINTS, M.H.A. Newman.

(67037-6) $6.95 THE Q U ~ A T I V E THEORY OF ORDINARY DIFFERENTIAL EQUATIONS: AN

INTRODUCTION, Fred Brauer and John A. Nohel. (6584G5) $8.95 FOURIER SERIES AND ORTHOGONAL FUNCTIONS, Harry F. Davis. (65973-9)

$1 1.95 TENSORS, DIFFERENTIAL FORMS, AND VARIATIONAL PRINCIPLES, David Lovelock

and Hanno Rund. (65840-6) $10.95 ESSENTIAL CALCULUS WITH APPLICATIONS, Richard A. Silverman. (66097-4)

TOPOLOGY, John G. Hocking and Gail S. Young. (656764) $10.95 NUMBER THEORY AND ITS HISTORY, Oystein Ore. (65620-9) $9.95 DISTRIBUTION THEORY AND TRANSFORM ANALYSIS, A.H. Zemanian. (654 79-6)

$11.95 INTRODUCTION TO NUMERICAL ANALYSIS, F. B. Hildebrand. (65363-3) $15.95 R m CHALLENGING PROBLEMS IN PROBABILITY WITH SOLUTIONS, Frederick

Mosteller. (653552) $4.95 HANDBOOK OF MATHEMATICAL FUNCTIONS, Milton Abrarnowitz and Irene A.

Stegun. (61272-4) $26.95 TENSOR ANALYSIS ON MANIFOLDS, Richard L. Bishop and Samuel I.

Coldberg. (64039-6) $8.95 VECTOR AND TENSOR ANALYSIS WITH APPLICATIONS, A.I. Borisenko and I.E.

Tarapov. (63833-2) $8.95 THE HISTORY OF THE CALCULUS AND ITS CONCEPTUAL DEVELOPMENT, Carl B.

Boyer. (60509-4) $9.95 PRINCIPLES OF STATISTICS, M.G. Bulrner. (63760-3) $8.95 THE THEORY OF SPINORS, f lie Cartan. (64070-1) $7.95 ADVANCED NUMBER THEORY, Harvey Cohn. (64023-X) $8.95 STATISTICS MANUAL, Edwin L. Crow, Francis Davis, and Margaret Maxfield.

(60599-X)

(continued on back flap)

Page 3: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

T3URIER SERIES Ceorgi I! Tolstov

Translated from the Russian by

Richard A. Silverman Richarc! A. Silvcrmatl's scrics of ti-allslarions of otrtslanding Rns- sian textbooks itn<l motlographs is well-known to pcople in tllc fielcls of mat1 tct~la tics, 1111 ysics and cttginccl-ing. Thc present book is another cxcellcrit tcxt fro111 [his sctmies, a valual~lc ntlditiorl to the English -1;ulguage li lernture o ~ l Fourier series.

This eclition is organizer1 illto nine well-clefincd chaprcrs: Trigono- metric Fouricr Scrics, Orthogonal Systcnls. Convergence of Trigono- metric Fouricr Scrics, Trigono~~ietric Series wit11 Decreasing Coefficients, Opcrarions OII Fou~~icr Series. Sunlmatiori of Trigono- nletric Fourier Scrics, I)ottl~lo IToi~ricr Scrics ant! tile Fourier Integral, Besscl Fuuclio~is a11d IToi~rier-Resscl Series, ancl the Eigcnfunction ilfcthotl and its Appliratio~ls to Mathel~atica! Pllysics. Every chaptcr r~tovcs tlcal-ly frotlr lopic to topic ;111cl thcoren~ to thcorcm, with malip thcolrnl proofs givct~. A total of 107 prohlerrls will bc Foutld at lhc ciitls of thc chapters, inclucling many specially added to this El~glisll -1anglt age cdi tio~i, and answers arc givcn at t hc c t~d of the tcxt.

Richarcl Silverrnan's excetllcnt tt;tnslation makes this book rcadily acccssil~lc LO ~natf~cm:~t ic ia~~s attd tttath sttttlents, as well as workers and stticlents in the fields of pliysics ant1 cnginccring. Hc has also addciI :L hihliogra phy , contai ning suggestiolls for collatcraf and supp1emental.y reading.

Unabridged, slightly corrcctc(1 rcp~r hIicatio11 o f thc or-iginal (1962) edition. 107 problctils. r\~lswers lo prol~lc~l~s. Bi hliogrnpl~y. Index. xi + 3SGpp. 5% x 8G. Pnperl>oun~l.

Free Dwer Mathematics iind Science Catalog (59065-8) available upon request.

I S B N 0-48b-b3317-9

$11-95 IN USA

Page 4: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

(continued from front flap)

COMPUTABILITY AND UNSOLVABIL~TY, Martin Davis, (61471-9) $8.95 A s m o n c METHODS M ANALYSIS, N. G. de Bruijn. (64221-6) $7.95 PROBUMS IN GROUP THEORY, John D. Dixon. (615744) $7.95 THE ATH HE MA TICS OF CAMB OF STRATEGY, Melvin Dresher. (64216-X) $6.95 Asmonc EXPANSIONS, A ErdClyi. (603184) $4.95 COMPLEX VARIABLES: HARMONIC AM) ANALYTIC F~CTIONS, Francis J. Flanigan.

(61388-7) A HISTORY OF GREEK M A I ~ ~ T I C S , Sir Thomas Heath. (240734,24074-6)

Two-volume set $25.90 MATHEMATICAL FOUNDATIONS OF INFORMATION THEORY, A.I. Khinchin. (60434-9)

$5.95 INTRODUCTORY RFAL ANALYSIS, AN. Kolmogomv and S.V. Fomin. (61'226.0)

$10.95 SPECIAL Fuvmoris AND THEIR APPUCATIONS, N.N. Lebedev. (60624-4) $9.95 F ~ ~ ~ A M E N T A L CONCEPTS OF ALGEBRA, Bruce E. Meserve. (61470-0) $9.95 F ~ ~ ~ A M E N T A L CONCEPTS OF GEOMETRY, Bruce E. Meserve. (63415-9) TA~LE~ OF I N D ~ INTEGRALS, G. Petit Bois. (602257) $8.95 MATRICB AND TRANSFORMATIONS, Anthony J. Pettofrezzo. (63634-8) $6.95 P R O B A B ~ THEORY: A CONCISE COURSE, Y.A. Rozanov. (63544-9) $7.95 ORDINARY DIFFERENTIAL EQUATIONS AND STABILITY THEORY: AN INTRODUCTION,

David A. Shchez. (63828-6) $6.95 LINEAR ALGEBRA, Georgi E. Shilov. (63518-X) $11.95 PROBLEMS W P R O B A B I ~ THEORY, MATHEMATICAL STATISTICS AND THEORY OF

RANDOM F ~ ~ ~ o N s , AA. Sveshnikov. (637 1 7-4) $12.95 TENSOR CALCULUS, J.L. Synge and A. Schild. (63612-7) $9.95 CALCULUS OF VARIATIONS WlTH APPUCATDNS TO PHYSICS AND ENGINEERING,

Robert Weinstock. (63069-2) $8.95 INTRODUCTION TO VECTOR AND TENSOR ~ Y S E , Robert C. Wrede. (61879-X)

$10.95

Paperbound unless otherwise indicated. Prices subject to change with- out notice. Available at your book dealer or write for free catalogues to Dept. 23, Dover Publications, Inc., 31 East 2nd Street, Mineola, N.Y. 11501. PIease indicate field of interest. Each year Dover publishes over 200 books on fine art, music, crafts and needlework, antiques, lan- guages, literature, children's books, chess, cookery, nature, anthropol- ogy, science, mathematics, and other areas.

Manufactured in the U.S.A.

Page 5: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

FOURIER SERIES

Page 6: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

FOURIER SERIES

GEORGI P. TOLSTOV Professor of Mathematics Moscow State U ' e r s i t y

nansIuted from the Russian by

Richard A. Silverman

DOVER PUBLICATIONS, INC.

NEW YORK

Page 7: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Copyright @ 1962 by Richard A. Silverman. All rights reserved under Pan American and

International Copyright Conventions.

Published in Canada by General Publishing Com- pany, Ltd., SO Lesmill Road, Don Mills, Toronto, Ontario.

This Dover edition, first published in 1976, is an unabridged republication, with slight corrections, of the work originally published by Prentice-Hall, Inc., Englewood Cliffs, New Jersey in 1962.

Intmrtional Standard Book Number: 0-486-63317-9 Library of Congress Catalog Card Number: 75-41883

Manufactured in the United States of America Dover Publications, Inc.

180 Variclt Street Netv York, N.Y. 10014

Page 8: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

AUTHOR'S PREFACE

My book on Fourier series, which originally appeared in Russian, has already been translated into Chinese, German, Polish and Rumanian. I am very grateful to Dr. R. A. Silverman and to the RenticeHall Publishing Company for undertaking to prepare and publish an English version of the second Russian edition. It would be most gratifying to me if the book were to serve the needs of American readers. I would like to thank V. Y. Kozlov, L. A. Tumarkin and A. I. Plesner for the helpful advice they gave me while I was writing this book.

a. P. T.

TRANSLATOR'S PREFACE

The present volume is the second in a new series of translations of outstanding Russian textbooks and monographs in the fields of mathematics, physics and engineering, under my editorship. It is hoped that Professor Tolstov's book will constitute a valuable addi- tion to the English-language literature on Fourier series.

v

Page 9: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

vi TRANSLATOR'S PREFACE

The following two changes, made with Professor Tolstov's consent, are worth mentioning:

1. To enhance the value of the English-language edition, a large number of extra problems have been added by myself and Professor Allen L. Shields of the University of Michigan* We have consulted a variety of sources, in particular, A ~oliection of P r ~ ~ in Mathematical Physim by N. N . Lebedev, I. Pa Skals- kaya, and Y. S. Uflyand (Moscow, 1955), from which most of the problems appearing at the end of Chapter 9 have been taken. 2. To keep the number of cross references to a minimum, four chapters (8 and 9, 10 and 11) of the Russian original have been combined to make two chapters (8 and 9) of the present edition.

I have also added a Bibliography, containing suggestions for collateral and supplementary reading. Finally, it should be noted that sections marked with asterisks contain material of a more advanced nature, which can be amitted without loss of continuity.

R* A. S.

Page 10: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

CONTENTS

1 TRIGONOMETRIC FOURIER SERIES Page 1. 1 : Periodic Functions, 1. 2 : Harmonics, 3. 3 : Trigonometric Polynomials and Series, 6. 4: A More Precise Terminology. Integrability. Series of Functions, 8. 5: The Basic Trigo- nometric System. The Orthogonality of Sines and Cosines, 10. 6: Fourier Series for Functions of Period 2sr, 12. 7: Fourier Series for Functions Defined on an Interval of Length 275 15. 8 : Right-hand and Left-hand Limits. Jump Discontinuities, 17. 9: Smooth and Piecewise Smooth Functions, 18. 10: A Criterion for the Convergence of Fourier Series, 19. 11 : Even and Odd Functions, 21. 12: Cosine and Sine Series, 22. 13: Examples of Expansions in Fourier Series, 24. 14: The Complex Form of a Fourier Series, 32. 1 5 : Functions of Period 21, 35. Problems, 38.

ORTHOGONAL SYSTEMS Page 41. 1 : Definitions, 41. 2: Fourier Series with Respect to an Orthogonal System, 42. 3: Some Simple Orthogonal Systems, 44. 4: Square Inte- grable Functions. The Schwarz Inequality, 50. 5: The Mean Square Error and its Minimum, 51. 6: Bed's Inequality, 53. 7 : Complete Systems. Convergence in the Mean, 54. 8 : Important Properties of Complete Systems, 57. 9: A Criterion for the Completeness of a System, 58. *lo: The Vector Analogy, 60. Problems, 63.

CONVERGENCE OF TRIGONOMETRIC FOURIER 3 SERIES Page 66. 1 : A Consequence of -1's Inequality. 66. 2: The Limit as n+oo of the Trigonometric Integrals

J b f ( ) a cos nx dx and P f ( x ) sin nr dx, 67. 3 : Formula for

the Sum of Cosines. Auxiliary Integrals, 71. 4: The Integral Formula for the Mia1 Sum of a Fourier Series, 72. 5 : Right-Handand LeA-HandDerivatives, 73. 6: A Sufficient Condition for Convergence of a Fourier Series at a Con- tinuity Point, 75. 7: A Sufficient Condition for Convergence

vl i

Page 11: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

of a Fourier Series at a Point of Discontinuity, 77. 8: Generalization of the SuWent Conditions Proved in Secs. 6 and 7, 78. 9: Convergence of the Fourier Series of a Piecewise Smooth Function (Continuous or Discontinuous), 79. 10: Absolute and Uniform Convergence of the Fourier Series of a Continuous, Piecewise Smooth Function of Period 2rr, 80. 1 1 : Uniform Convergence of the Fourier Series of a Continuous Function of Period 2% with an Absolutely Integrable Derivative, 82. 12: Gendzation of the Results of Sec. 11, 85. 13: The Localization Principle, 90. 14: Examples of Fourier Series Expansions of Unbounded Func- tions, 91. 15 : A Remark Concerning Functions of Period 21,94. Problems, 94.

4 TRIGONOMETRIC SERIES WITH DECREASING COEFFICIENTS Page 97. 1: Abel's Lemma, 97. 2: Fomula for the Sum of Sines. Auxiliary Inequalities, 98. 3: Convergeme of Trigonometric Series with Monotonically Decreasing Coefficients, 100. *4: Some Comqmces of the Theorems of Sec. 3, 103. 5: Applications of Functions of a Complex Variable to the Evaluation of Certain Triguno- metric Series, 105. 6: A Stronger Form of the M t s of Sec. 5, 108. Problems, 112.

OPERATIONS ON FOURIER SERIES Page 115. 1: Approximation of Functions by Trigonometric Polynomials, 1 15. 2 : Completeness of the Trigonometric System, 1 17. 3 : Parseval's Theorem. The Most Important Conseqwnces of the Completeness of the Trigonometric System, 119. *4: Approximation of Functions by Polynomials, 120. 5: Addition and Subtraction of Fourier Series, Multiplication of a Fourier Series by a Number, 122. *6: Products of Fourier Series, 123. 7: Integration of Fourier Series, 125. 8: Differentiation of Fourier Series. The Case of a Con- tinuous Function of Period 2n, 129. $9: Differentiation of Fourier Series. The Case of a Function Dew on the Interval [- n, n], 132. * 10: Differentiation of Fourier Series. The Case of a Function Defined on the Interval [0, IT], 137. 11 : Improving the Convergeme of Fourier Series, 144. 12: A List of Trigonometric Expansions, 147. 13 : Approximate Calculation of Fourier Coetficients, 150. Boblems, 152.

6 SUMMATION OF TRIGONOMETRIC FOURIER SERIES Page 155. 1 : Statement of the Problem, 155. 2 : The Method of Arithmetic Means, 156. 3: The Integral

Page 12: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Formula for the Arithmetic Mean of the Partial Sums of a Fourier Series, 157. 4: Summation of Fourier Series by the Method of Arithmetic Means, 158. 5 : Abel's Method of Summation, 162. 6 : Poisson's Kernel, 1 63. 7 : Application of Abel's Method to the Summation of Fourier Series, 164. Problems, 1 70.

DOUBLE FOURIER SERIES. THE FOURIER INTEGRAL Page 173. 1 : Orthogonal Systems in Two Variables, 173. 2: The Basic Trigonometric System in Two Variables. Double Trigonometric Fourier Series, 1 75. 3 : The Integral Formula for the Partial Sums of a Double Trigonometric Fourier Series. A Convergence Criterion, 178. 4 : Double Fourier Series for a Function with Different Periods in x and y, 180. 5: The Fourier Integral as a Limiting Case of the Fourier Series, 180. 6: Improper Integrals Depending on a Parameter, 182. 7 : Two Lemmas, 185. 8: Roof of the Fourier Integral Theorem, 188. 9: Different Forms of the Fourier Integral Theorem, 189. * 10: The Fourier Transform, 190. * 11 : The Spectral Function, 193. Problems, 195.

BESSEL FUNCTIONS AND FOURIER-BESSEL SERIES Page 197. 1 : Bessel's Equation, 197. 2 : Bessel Functions of The First Kind of Nonnegative Order, 198. 3 : The Gamma Function, 201. 4: Bessel Functions of the First Kind of Negative Order, 202. 5: The General Solution of Bessel's Equation, 203. 6: Bessel Functions of the Second Kind, 204. 7: Relations between Bessel Functions of Different Orders, 205. 8: Bessel Functions of the F i t Kind of Half-Integral Order, 207. 9 : Asymptotic Formulas for the Bessel Functions 208. 10: Zeros of the Bessel Functions and Related Func- tions, 21 3. 1 1 : Parametric Form of Bessel's Equation, 21 5. 12 : Orthogonality of the Functions Jp(h), 216. 13: Evalua- tion of the Integral J' x.lp2(hx) d*, 21 8. * 14: Bounds for the

0

Integral J ' x ~ , ~ h r ) &, 219. 1 5 : Definition of Fourier-Bessel 0

Series, 220. 16: Criteria for the Convergence of Fourier- Bessel Series, 221. $17: -1's Inequality and its Conse- quences, 223. + 18 : The Order of Magnitude of the Coeffi- cients which Guarantees Uniform Convergence of a Fourier- Bessel Series, 225. $19: The Order of Magnitude of the Fourier-Bessel Coefficients of a Twice Differentiable Function, 228. $20: The Order of Magnitude of the Fourier-Bessel Coefficients of a Function Which is Differentiable Several

Page 13: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Times, 231. *21: Term by Term Differentiation of Fourier- Ressel Series, 234. 22: Fourier-Bessel Series of the Second '1[Lpe, 237. *23 : Extension of the Results of Sccs. 17-21 to Fourier--1 Series of the Second Type, 239. 24: Fourier- Besscl Expansions of Functions Dehed on the Interval [O, 11,241. Problems, 243.

THE EIGENFUNCI'ION METHOD AND ITS APPLICA- TIONS TO MATHEMATICAL PHYSICS Page 245. Part I: THEORY. 1: The Gist of the Method, 245. 2: The Usual Statement of the Boundary Value Problem, 250. 3: The Existence of Eigendues, 250. 4: EigRnfdons and Their Orthogonality, 251. 5: Sign of the Eigenvalues, 254. 6: Fourier Series with Respect to the Eigenhnctions, 255. 7 : Does the Eigenfunction Method Always Lead to a Solution of the Roblem?, 258. 8 : The Generalized Solution, 261. 9 : The Inhomogeneous Problem, 264. 10 : Supplementary Remarks, 266. Part 11: APPLICATIONS, 268. 11 : Equation of a Vibrating String, 268. 12: Free Vibrations of a String, 269. 13: Forced Vibrations of a String, 273. 14: Equation of the Longitudinal Vibrations of a Rod, 275. 15: Free Vibrations of a Rod, 277. 16: Forced Vibrations of a Rod, 280. 17 : Vibrations of a Rectmguh Membrane, 282. 18 : Radial Vibrations of a C i d a r Membrane, 288. 19: Vibrations of a Circular Membrane (General Case), 291. 20: Equation of Heat Flow in a Rod, 296. 21 : Heat Flow in a Rod with Ends Held at Zero Temperature, 297. 22: Heat Flow in a Rod with Ends Held at Constant Temperature, 299. 23: Heat Flow in a Rod Whose Ends are at Specified Variable Temperatures, 301. 24. Heat FIow in a Rod Whose Ends Exchange Heat Freely with the Surrounding Medium, 301. 25: Heat Flow in an Infinite Rod, 306. 26: Heat Flow in a Circular Cylinder Whose Surface is Insulated, 310. 27: Heat Flow in a Circular Cylinder Whose Surface Exc- Heat with the Surrounding Medium, 312. 28: Steady-State Heat Flow in a Circular Cyiinder, 3 13. Problems, 3 16.

ANSWERS TO PROBLEMS Page 319.

BIBLIOGRAPHY Page 331.

INDEX Page 333.

Page 14: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

FOURIER SERIES

Page 15: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

TRIGONOMETRIC FOURIER SERIES

I. Periodic Functions

A function f(x) is called periodic if there exists a constant T > 0 for which

for any x in the domain of definition of f(x). (It is understood that both x and x + T lie in this domain.) Such a constant T is called a period of the function f(x). The most f d a r periodic functions are sin x, cos x, tan x, etc. Periodic functions arise in many applications of mathematics to problems of physics and engineering. It is clear that the sum, difference, product, or quotient of two functions of period T is again a function of period T.

If we plot a periodic function y = f(x) on any closed interval a < x < a + T,-we can obtain the entire graph of f(x) by periodic repetition of the portion of the graph corresponding to a < x < a + T (see Fig. 1).

If T is a ptriod of the function f (x), then the numbers ZT, 3T, 4T, . . . are also periods. This follows immediately by inspecting the graph of a periodic functiob or from the series of equalities'

1 We suggest that the reader prove the validity not only of these equalities but also of the following equalities: ,

f(x) = f (x - T) = f(x - 2T) = f(x - 3T) = I

Page 16: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

2 TIUaONOIHETRICFO~SERIES CHAP. 1

which are obtained by repeated use of the condition (1.1). Thus, if T is a period, so is KT, where k is any positive integer, i.e., if a period exists, it is not unique.

Next, we note the following property of any function f (x) of period T:

Iff(,) is irtegrable on any Interval of length T, then it b htegrable on any other interval of the same length, and the vahie of the integral tr the same, i.e.,

for any a and b.

This property is an immediate consequence of the interpretation of an integral as an area. In fact, each integral (1.2) equals the area induded between the curve y = f(x), the x-axis and the ordinates drawn at the end points of the interval, where areas lying above the x-axis are regarded as positive and ar&s lying below the x-axis are regarded as negative. In the present case, the areas represented by the two integrals are the same, because of the periodicity of f(x) (sa Fig. 2).

H d t e r , when we say that a function f(x) of period T is integruble, we shall mean that it is integrable on an interval of length T. It follows from

Page 17: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 2 TRIGONOMETRIC FOURIER SERIES 3

the property just proved that f(x) is also integrable on any interval of finite length.

2. Harmonics

The simplest periodic function, and the one of greatest importance for the applications, is

y = A sin (on + p), where A, a , and cp are constants. This function is called a harmonic of amplitude I AI, (angular) frequency o, and initial phase 9. The period of such a harmonic is T = h/o, since for any x

The terms "amplitude, " "frequency, " and "initial p h a ~ " stem from the following mechanical problem involving the simplest kind of oscillatory motion, i.e., simple harmonic motion: Suppose that a point mass M, of mass m, moves along a straight line under the action of a restoring force F which is proportional to the distance of M from a fixed origin 0 and which is directed towards 0 (see Fig. 3). Regarding s as positive if M lies to the right of 0 and

negative if M lies to the left of 0, i.e., assigning the usual positive direction to the line, we find that F = -ks, where k > 0 is a constant of proportionality. Therefore

- where we have written o2 = k/m, so that o = .\/k/m.

It is easily verified that the solution of this differential equation is the function s = A sin (a t + cp), where A and cp are constants, which can be calculated from a knowledge of the position and velocity of the point M at the initial time t = 0. This function 8 is a harmonic, and in fact is a periodic function of time with period T = 27tlo. Thus, under the action of the

Page 18: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

4 T R I 0 0 N ~ C F O U R I E R S e R l E S CHAP. 1

restoring force F, the point M undergoes oscillatory motion. The amplitude IAI is the maximum deviation of the point M from 0, and the quantity 1/T is the number of oscillations in an interval containing 27c units of time (e.g., seconds). This explains the term "frequency'*. The quantity 0 is the initial phase and characterizes the initial position of the point, since for t = 0 we have so = sin cp.

We now examine the appearance of the curve y = A sin (ax + 9). We assume that o > 0, since otherwise sin (- o x + 9) is merely replaced by - sin (ax - 9). The simplest case is obtained when A = 1, w = 1, cp = 0; this gives the familiar slrre curve y = sin x [see Fig. 4(a)]. For A = 1, w = 1, cp = z/2, we obtain the c&e cww y = cos x, whose graph is the same as that of y = sin x shifted to the left by an amount x/2.

Next, consider the harmonic y = sinox, and set o x = Z, thereby obtaining y = sin x, an ordinary sine curve. Thus, the graph of y = sin ox is obtained by deforming the graph of a sine curve: This deformation

Page 19: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 2 TRIGONOMETRIC FOURIER SERIES 5

reduces to a uniform compression along the x-axis by a factor w if o > 1, and to a uniform expansion along the x-axis by a factor l/o if o < 1. Figure 4(b) shows the harmonic y = sin 3x, of period T = 243.

Now, consider the harmonic y = sin (ox + cp), and set ox + cp = oz, so that x = z - cp/o. We already know the graph of sin oz. Therefore, the graph of y = sin (ox + cp) is obtained by shifting the graph of y = sin ax along the x-axis by the amount -cp/o. Figure 4(c) represents the harmonic

y = sin (3x + ;) with period 2 4 3 and initial phase 4 3 .

Finally, the graph of the harmonic y = A sin (ox + p) is obtained from that of the harmonic y = sin (ox + cp) by multiplying all ordinates by the number A. Figure 4(d) shows the harmonic

y = 2 sin (3x + $)* These results may be summarized as follows:

The graph of the harmonic y = A sin (wx + Q) is obtained from the graph of the fmiiiar sine curve by unffom compression (or expansion) along the coordinate axes plus a shift along the x-a&.

Using a well-known formula from trigonometry, we write

A sin (wc + cp) = A(cos o x sin p + sin ox cos 9).

Then, setting

a = A sin cp, b = A cos p, (2.1)

we convince ourselves that every harmonic can be represented in the form

a cos wx + b sin ax. (2.2)

Conversely, every function of the form (2.2) is a harmonic. To prove this, it is sufficient to solve (2.1) for A and B. The result is

Q a A = .\/a2 + b2, sin cp = - =

b cosp = - = b A 1 / 4 2 + b2' A vPTT~'

from which cp is easily found. From now on, we shall write harmonics in the form (2.2). For example,

for the harmonic shown in Fig. 4(d), this form is

2 sin 3x + - = d3 cos 3x + sin 3x ( 3

Page 20: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

6 TRIOONOM~~RIC mmm SER~ES CHAP. 1

It will also be convenient to explicitly introduce the period T in (2.2). If we set T = 2Z, then, since T = 2x/w, we have

and therefore, the harmonic with period T = 22 can be written as

7t;Y XX acos- + bsin- I I

3. Trigonometric Polynomials and Series

Given the period T = 21, consider the harmonics

xkx xkx ak cos - + bk sin - I 1 (k = 1,2,. . .)

with frequencies wk = ~ k / l and periods Tk = 2 h = Z/k. Since

T= 21 = kTk,

the number T = 21 is simultaneously a period of all the harmonics (3.1), for an integral multiple of a period is again a period (8ee Sec. 1). Therefore, every sum of the form

xkx sj,(x) = A + 2 (ak cos - + bk sin

k= 1 1

where A is a constant, is a function of period 21, since it is a sum of functions of period ZI. (The addition of a constant obviously does not destroy periodicity; in fact, a constant can be regarded as a function for which any number is a period.) The function sn(x) is called a trigonometric polynomial of order n (and period 21).

Even though it is a sum of various harmonics, a trigonometric poly- nomial in general represents a function of a much more complicated nature than a simple harmonic. By suitably choosing the constants A, al, bl, a2, b2,. . . we can form functions y = sn(x) with graphs quite unlike the smooth and symmetric graph of a simple harmonic. For example, Fig. 5 shows the trigonometric polynomial

y = sin x + .) sin 2x + 4 sin 3x.

The infiite trigonometric series

xkx A + 2 (a. cos - + bk sin - k=1 1 ") 1

Page 21: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 3 TRIGONOMETRIC FOURIER SERIES 7

(if it converges) also represents a function of period 22. The naturs of func- tions which are sums of such infinite trigonometric series is even more diverse. Thus, the following question arises naturally: Can any given

function of period T = 22 be represented as the sum of a trigonometric series? We shall see later that such a representation is in fact possible for a very wide class of functions.

For the time being, suppose that f(x) belongs to this class. This means that f(x) can be expanded as a sum of harmonics, i r , as a sum of functions with a very simple structure. The graph of the function y = f(x) is obtained as a "superposition" of the graphs of these harmonics. Thus, to give a mechanical interpretation, we can rep-sent a complicated oscillatory motion f(x) as a sum of individual oscillations which are particularly simple. How- ever, one must not imagine that trigonometric series are applicable only to oscillation phenomena. This is far from being the case. In fact, the concept of a trigonometric series is also very useful in studying many phenomena of a quite different nature.

If

nkx

k= 1 2 f(x9 = A + 5 (ak coa - + bk sin -

then, setting zx/l= t or x = t l 'q we find that

QO

90) = f((II x = A + + (ak cos kt + bk sin kt), k= 1

where the harmonics in this series all have period 2 ~ . This means that if a function f(x) of period 21 has the expansion (3.2), then the function cp(t) = f(tl/se) is of period 27c and has the expansion (3.3). Obviously, the converse is also true, i.e., if a function cp(t) of period 2x has the expansion (3.3), then the function f(x) = cp(rur/l) is of period 21 and has the expansion (3.2).

Page 22: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

8 T R I O O N O ~ C FOURIER SERIES CHAP. 1

Thus, it is enough to know how to solve the problem of expansion in trigono- metric series for functions of the "standard" period 27~. Moreover, in this case, the series has a simpler appearance. Therefore, we shall develop the theory for series of the form (3.3), and only the final results will be converted to the " language " of the general series (3.2).

4. A More Precise Terminology. Integrability. Series of Functions

We now introduce a more precise terminology and recall some facts from differential and integral calculus. When we say that f (x) is integrable on the interval [a, b], we mean that the integral

(which may be improper) exists in the elementary sense. Thus, our inte- grable functions f(x) will always be either continuous or have a finite number of points of discontinuity in the interval [a, b], at - which the function can be either bounded or unbounded.

In courses on integral calculus, it is proved that if a function has a finite number of discontinuities, then if the integral

exists, so does the integral (4.1). (The converse is not always true.) In this case, the function f(x) is said to be absolutely integrable. If f(x) is absolutely integrable and cp(x) is a bounded integrable function, then the product f (x)q(x) is absolutely integrable. The following rule for integration by parts holds :

Let f(x) and p(x) be continuow on [a, b], but perhaps non-dtrerentiable at a finite munber of points. Then, iff (x) and cpf(x) are absolutely inte- grable,z we have

Another familiar result is the fact that if the functions fi(x),f2(x), . . .,&(x) are integrable on [a, b], then their sum is also integrable, and

2 Instead of absolute integrability of both derivatives, we can weaken this requirement to absolute integrability of just one of the derivatives. However, the stronger form of the requirement is sufficient for what follows.

Page 23: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 4 TRIGONOMETRIC FOURIER S E R ~ 9

We now consider an infinte series of functions

Such a series is said to be convergent for a given value of x if its partial sums

have a finite limit

The quantity s(x) is said to be the mm of the series, and is obviously a function of x. If the series converges for all x in the interval [a, b], then its sum s(x) is defined on the whole interval [a, b].

We now ask whether the formula (4.3) can be extended to the case of a convergent series of functions which are integrable on the interval [a, b], i.e., is the formula

valid? In other words, can the series be integrated term by term? It turns out that (4.5) is not always valid, if for no other reason than that a series of integrable or even continuous functions may not even have an integrable sum. A similar problem arises in connection with the possibility of term by term differentiation of series. We now single out an important class of series of functions to which these operations can be applied.

The series (4.4) is said to be uniformly convergent on the interval [a, b] if for any positive number r, there exists a number N such that the inequality

holds for all n 3 N and for aN x in the interval [a, b]. Thus, if we examine the graph of the sum of the series s(x) and of the partial sum s.(x), uniform convergence means that for all sufficiently largeindices n and for all x, the curve representing s(x) and the curve representing sn(x) are less than c apart, where c is any preassigned number, so that the two curves are unifomly) close (see Fig. 6).

Not every series which converges on an interval [a, b] converges uniformly there. The following is a very useful and simple test for the uniform con- vergence of a series of functions (Weierstrass' M-test) :

3 I.e., for all x in [a, 61.

Page 24: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

10 TRIGONOMETRIC FOURIER SERIES CHAP. 1

lf the series of positive numbers MI + M2 +***+Mk + * * a

converges and i f for any x in the interval [a, b] we have (h(x)l < Mk jiom a certain k on, then the series (4.3) converges uniformly (and absolutely) on [a, b].

The following important theorems are valid :

THEOREM 1. If the terms of the series (4.4) are continuous on [a, b] and if the series is un formly convergent on [a, b I , then

a) The sum of the series is continuous; b) lk sum can be integrated term by term, i.e., (4.5) hol&.

THEOREM 2 If the series (4.4) converges, flits terms are dlrerentiable and d f the series

Ir un formly convergent on [a, b], t h

Le., the series (4.4) can be dtrerentiated term by term.4

5. The Basic Trigonometric System. The Orthogonality of Sines and Cosines

By the basic trigonometric system we mean the system of functions

1, cos x, sin x, cos 2x, sin 2x, . . . , cos nx, sin nx, . . . (5.1)

4 In courses on analysis, it is usually assumed also that the derivatives are continuous, in order to simplify the proof.

Page 25: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 5 TRIGONoMETRIC FOURIER SERIES f f

All these functions have the common period 27c (although cos nx and sin nx also have the smaller period 2412). We now prove some auxiliary formulas.

For any integer n # 0, we have

n I-, x- -n

cos nx X'n r n sin nx dx = [- ] = 0,

X 3 -n

and

rr cos2nx dx = 1, 1 + cos 2nx 2 dx = n,

Using the familiar trigonometric formulas

cos a cos p = +[COS (a + p) + cos (a - (j)], sin a sin p = + [cos (a - p) - cos (a + p)]

we find that

r" cos nx cos mx dx

cos (n + m)x + cos (n - *)XI dx = 0, = +rn1

rn sin nx sin mx dx

= + $I [cos (n - m)x - cos (n + m)xl tix = o -n

for any integers n and m (n # m). Finally, using the formula

sin a cos (j = +[sin (a + p) + sin (a - (j)], we find that

C" sin nx cos mx dr

= [sin (n + m)x + sin (n - m)x] dx = 0 -i

for any'n and m. The formulas (5.2). (5.4). and (5.5) show that the integral

Page 26: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

12 TRIOoNOblETRIC FOURIER 8ERIES CHAP. 1

over the interval [- x, rr] of the product of any two diflerent functions of the system (5.1) vanishes.

We shall agree to call two functions cp(x) and +(x) orthogonds on the interval [a, b] if

With this definition, we can say that the functions of the system (5.1) are pairwise orthogonal on the interval [-n, rr], or more briefly, that the system (5.1) 6 orthogonal on [- rr, IT].

As we know, the integral of a periodic function is the same over any interval whose length equals the period (see Sec. 1). Therefore, the formulas (5.2) through (5.5) are valid not only for the interval [-IT, x] but also for any interval [a, a + 2x1, i.e., the system (5.1) is orthogonal on every such interval.

6. Fourier Series for Functions of Period IT

Suppose the function f (x) of period 2% has the expansion a0

+ 2 (a, cos kx + bk sin kx), f(x) = k= 1

where, to simplify the subsequent formulas, we denote the constant term by a0/2. We now pose the problem of determining the coefficients ao, ak and b (k = 1.2,. . .) from a knowledge of f(x). To do this, we make the following awumption : It is assumed that the series (6.1), and the series to be written presently, can be integrated term by term, i.e., it is assumed that for all these series the integral of the sum equals the sum of the integrals. pt is thereby dm assumed that the function f (x) is integrable.] Then, integrating (6.1) from -z to x, we obtain

By (5.2), all the integrals in the sum vanish, so that

Next, we multiply both sides of (6.1) by cos n* and integrate the result from - z to rr, as before, obtaining

5 In m, the word ortIbogonuIity connotes perpendicularity. One must not think that the concept of orthogonality of two functions corresponds to anything like perpendi- cularity of their graphs, despite the fact that this concept is related to a suitably generalized notion of perpendicularity. In this w r d , see Ch. 2, Sec. 10.

Page 27: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Lx f(x) cos nxdx = y ' = = c o s ~ e d * -

By (5.2), the first integral on the right vanishes. Since the functions of the system (5.1) are painvise orthogonal, all the integrals in the sum also vanish, except one. The only integral that remains is the coefficient of an:

[see (5.3)]. Thus we have

Similarly, we find that

r= f (x) sin nx dx = bnZ.

It follows from (6.2) to (6.4) that

1 bn = - rnf (x) sinnxdx (n = 1.2,. ..). X

Thus, finally, if f(x) is integrable and can be expanded in a trigonometric series, and if this series and the series obtained from it by multiplying by cos nx and sin nx (n = 1.2, . . .) can be integrated term by term, then the coefficients an and bn are given by the formulas (6.5).

Now, suppose we are given an integrable function f(x) of period 2x, and we wish to represent f(x) as the sum of a trigonometric series. If such a representation is possible at all (and if the requirement of term by term integrabiity is satisfied), then by what has been said, the coefficients an and bn must be given by (6.5). Therefore, in looking for a trigonometric series whose sum is a given function f(x), it is natural to examine first the series whose coefficients are given by (6.5), and to see whether this series has the required properties. As we shall see later, this will be the case for a large class of functions.

The cafficients 4, and bn calculated by the formulas (6.5) called the F d e r coeficients of the function f(x), and the trigonometric series with

Page 28: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

14 TRIGONOMETRIC FOURIER SERIES CHAP. 1

these cdcients is caned the Fourier series off (x). Incidentally, we note that the formulas (6.5) involve integrating a function of period 2x. There- fore, the interval of integration [-=, rr] can be replaced by any other interval of length 21t (see Sec. I), so that together with the formulas (6.5), we have

The above considerations make it natural to devote special attention to Fourier series. It we form the Fourier series of a function f(x) without deciding in advance whether it converges to f(x), we write

00

f ( ~ ) - + (an COS nx + 4 sin nx). n=1

This notation means only that the Fourier series written on the right corre- sponds to the function f(x). The sign - can be replaced by the sign = only if we succeed in proving that the series converges and that its sum equals f(x). A simple consequence of these considerations is the following theorem, which is quite useful:

THEOREM 1. a f i i o n f (x) of period 21t cun be expanded in a trigonometric series which converges unformly on the whole real axis,6 then this series b the Fourier series of f(x).

Proo$ Suppose that f(x) satisfies (6.1), where the series is uni- formly convergent. By Theorem 1 of Sec. 4, f(x) is continuous and term by term integration of the series is possible. This gives the formula (6.2). Next, we consider the equality

f(x) cos nx = cos nx 00

+ 2 (ak cos kx cos nx + bk sin kx cos nx), (6.7) k= 1

and show that the series on the right is uniformly convergent. Set m

@ + 2 (ak cos kx + bk sin kx), sm(x) = T k= 1

and let E be an arbitrary positive number. If the. series (6.1) converges uniformly, then there exists a number N such that

6 By the periodicity of f(x) we can require uniform convergence on [-z, TC], rather than on the whole real axis.

Page 29: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 6 TRIGONOMETRIC FOURIER SERIES I 5

for all m N. The product s,,,(x) cos nx is obviously the mth partial sum of the series (6.7). Then, the inequality

I f(x) cos nx - s,,,(x) cos nxl = If (x) - sm(x)l lcos nxl r E,

which holds for all m 2 N, implies the uniform convergence of the series (6.7). It follows that this series can be integrated term by term, and the result of the integration is the formula (6.3). Similarly, we prove the formula (6.4). Thus, finally, the formulas (6.5) hold for the coefficients a, and b,, which means that (6.1) is the Fourier series of f (XI*

The modern theory of Fourier series allows us to prove the following more general result, whose proof we cannot give because of its complexity:

THEOREM 2. If an absolutely integrable function f(x) of period 2rr can be expanded in a trigonometric series which converges to f(x) every- where, except possibly at a finite number of points (within one period), then this series is the Fourier series of f(x).

This thtorem confirms the assertion made above, that in looking for a trigonometric series which has a given function f(x) as its sum, we should first consider the Fourier series of f(x).

7. Fourier Series for Functions Defined on an Interval of Length 2n

A problem which arises quite often in the applications is that of expanding a function f(x) in trigonometric series, when f(x) is defined only on the interval [-n, n]. In this case, nothing at all is said about the periodicity of f(x). Nevertheless, this does not prevent us from writing the Fourier series of f(x), since the formulas (6.5) involve only the interval [-n, A]. Moreover, f(x) can be extended by periodicity from [-x, x] onto the whole x-axis. This leads to a periodic function which coincides with f(x) on [-A, x] and which has a Fourier series identical with that of f(x). In fact, if the Fourier series of f(x) turns out to converge to f(x), then, sin- it is a periodic function, the sum of this Fourier series automatically gives us the required periodic extension of f(x) from [-n, rs] onto the whole x-axis.

Thus, it does not matter whether we talk about the Fourier series of a function defined on [-n, XI, or whether we talk about the Fourier series of the function obtained from f(x) by periodic extension along the x-axis. This implies that it is sufficient to formulate the tests for convergence of Fourier series for the case of periodic functions.

In connection with the problem of extending f(x) by periodicity from the interval [-n, A] onto the whole x-axis, the following remarks are in order: If f(-x) = f (~ ) , there is no difficulty in making the extension, since in this

Page 30: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

I6 TRIOONOMETRIC FOURIER SERIES CHAP. 1

case, iff(x) is continuous on [-x, x], its extension will be continuous on the whole x-axis [see Fig. 7(a) J. However, iff (-x) # f (x), we cannot accom- plish the required extension without changing the values off (-x) and f (x), since the periodicity requires that f(-rr) and f(n) coincide. This difficulty can be avoided in two ways: (1) We ccin completely avoid considering the values off(x) at x = -x and x = n, thereby making the function undefined at these points and hence making the periodic extension of f(x) undefined at the points x = (2k + lh k = 0, f 1, f 2, . . . ; (2) We can suitably modify the values of the function f(x) at x = - x and x = x by making these values equal. It is important to note that in both cases, the Fourier coefficients will have the same values as bdore, since changing the values of a function at a finite number of points, or even failing to define it at a finite number of points, cannot affect the value of an integral, in particular, the values of the integrals (6.5) defining the Fourier coefficients. Thus, whether or not we carry out the indicated modification of the function f(x), its Fourier series remains unchanged.

It should be observed that if f(-x) # f(x) and iff (x) is continuous on the interval [-n, x], then the periodic extension of f(x) onto the whole x-axis will have discontinuities at all the points x = (2k + I)IT, k = 0, f I, f 2, . . . , no matter how we change the values of the function at x = -n and x = x [see Fig. 7(b)]. The problem of finding the values to which the Fourier series of f(x) may be expected to converge at x = fx3 when f(-x) # f(x), is a special one, and will be solved later.

Finally, suppose that f(x) is defined on an arbitrary interval [a, a + 2x1 of length 2 ~ , and that it is required to expand .(x) in a trigonometric series.

Page 31: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 7 TRIGONOMETRIC FOURIER SERIES 1 7

As before, we arrive at the conclusion that it does not matter whether we talk about the Fourier series of f(x) or about the Fourier series of the function obtained from f(x) by extending it periodically onto the whole x-axis. If f(x) is continuous on the interval [a, a + 2x1 but f(a) # f(a + 24, we obtain an extension which is discontinuous at the points x = a + 2ksr (k = 0, * I , *2 ,... ).

8. Right-Hand and Left-Hand Limits. Jump Discontinuities

We introduce the notation

lim f (x) = f (xo - 01, am f (x) = f (xo + O), x+ xo x<xo

x-xo x>xo

provided these limits exist and are finite.' The first of these limits is called the lefr-hand limit of f(x) at the point xo, and the second is called the right- h d limit of f(x) at xo. These limits both exist at points of continuity (by the very definition of continuity), and we have

at continuity points. If xo is a point of discontinuity of the function f(x), then the right-hand

and left-hand limits (either or both of them) may exist in some cases and fail to exist in others. If both limits exist, we say that the point xo is a point of discontinuity of the first kind, or simply, a point of jump discontinuity. If at least one of these limits does not exist, then the point xo is called a point of discontinuity of the second kind. We shall be particularly interested in jump discontinuities. If xo is such a point, then the quantity

is called thejump of the function f(x) at xo. The following example illustrates this situation. Suppose that

-x3 for x < 1,

with the graph shown in Fig. 8. The value of the function at x = 1 is indicated by the little circle. At x = 1, the left-hand and right-hand limits are obviously

7 If xo = 0, we do not write f (0 + 0) and f (0 - O), but simply f (+ 0) and f (- 0).

Page 32: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

18 TRI~ON~METRIC FOURIER SERIES CHAP. 1

Therefore, the jump of the function at x = 1 is

8 = f(l + 0) - f(l - 0) = 2,

which is in complete agreement with the intuitive idea of a jump (see Fig. 8).

If f(x) is a function which is continuous on the interval [-x, n], then if f(-rr) # f(rr), jump discontinuities appear in making the periodic extension of f(x) from [ -rr , x] onto the whole x-axis [see Fig. 7(b)], and all the jump discontinuities are equal to the number

9. Smooth and Piecewise Smooth Functions

The function f(x) is said to be smooth on the interval [u, b] if it has a continuous derivative on [a, b]. In geometrical language, this means that the direction of the tangent changes continuously, without jumps, as it moves along the curve y = f(x) [see Fig. 9(a)]. Thus, the graph of a smooth function is a smooth curve without any "corners. "8

The function f(x) is said to be piecewise smooth on the interval [a, b] if either f(x) and its derivative are both continuous on [a, b], or they have only a finite number of jump discontinuities on [a, b]. It is easy to see that the graph of a piecewise smooth function is either a continuous curve or a dis- continuous curve which can have a finite number of corners (at which the derivative has jumps). As we approach any discontinuity or corner (from one side or the other), the direction of the tangent approaches a definite limiting position, since the derivative can have only jump discontinuities.

8 "Corner" = Russian "ymosarr ~orlxa," a point at which the curve has two distinct tangents. (7kanslator)

Page 33: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Figures 9(b) and 9(c) illustrate the graphs of continuous and discon- tinuous piecewise smooth functions. From now on, we shall regard smooth functions as a special case of piecewise smooth functions.

A continuous or discontinuous function f(x) which is defined on the whole x-axis, is said to be piecewise smooth if it is piecewise smooth on every interval of finite length. In particular, this concept applies to periodic functions. Every piecewise smooth function f(x) [whether continuous or discontinuous] is bounded and has a bounded derivative everywhere, except at its corners and points of discontinuity [at all these points, f'(x) does not exist].

10. A Crlterlon for the Convergence of Fourier Series

We now give a more useful criterion for the convergence of a Fourier series, deferring the proof of this criterion until Ch. 3:

The Fourier series of a piecewise smooth (continuous or discontinuous) function f(x) of period 2x converges for all values of x. The sum of the series equals f(x) at every point of continuity and equals the number

the arithmetic mean of the right-hand and left-hand limits, at every point of discontinuity (see Fig. 10). Iff (x) is continuous everywhere, then the series converges absolutely and uniformly.

Page 34: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

20 T R I O O N O ~ C FOURIER SERIES CHAP. 1

Suppose the function f(x) is defined only on [-x, z], and is pieoewise smooth on the interval [ox, z] and continuous at its end points. As noted in Sec. 7, the Fourier series of f(x) coincides with the Fourier series of the

function which is the periodic extension of f(x) onto the whole x-axis. But in this case, such an extension obviously leads to a function f(x) which is piecewise smooth on the whole x-axis. Therefore, the criterion just formu- lated implies that the Fourier series of f(x) will converge everywhere. In particular, the series will converge on the original interval [-x, x]. In fact, for - x < x < rr, the series will converge to f(x) at the points of continuity and to the value

at the points of discontinuity. But what will happen at the end points of the interval [ - x, x] ?

At the end points, two cases are possible:

1 f - ) = f ) . In this case, the periodic extension obviously leads to a function which is continuous at the points & x [and also at all the points x = (2k + l)z, k = 0, f 1, f 2,. . .I. Therefore, by our criterion, the Fourier series will also converge to f(x) at the end points of [-x, x].

2) f x ) # f ) . In this case, the periodic extension leads to a function which is discontinuous at the points fx [and also at the points x = (2 + I)%, k = 0, f 1, f 2,. . .I, where for the extension of f(x) we obviously have

(see Fig. 11). Therefore, at x = -x, and x = I, the Fourier series will converge to the values

Page 35: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

TRIGONOMETRIC FOURIER SERIES 2 1

Thus, the Fourier series of a function f(x) defined on the interval [-IT, n] and continuous at x = f x behaves at the points x = *x just as it does at the other points of continuity, provided that f(-IC) = ~(Ic). However, if

f(-n) # f(x), the series obviously cannot converge to f (x ) at x = x, and in this case, it is meaningful to pose the problem of expanding f(x) in Fourier series only for -A < x < x and not for - x 6 x 6 n. A similar remark can be made concerning the Fourier series of a function specified in an interval of the type [a, a + b], whexe a is any number,

In solving any concrete problem, if the reader draws a graph of the periodic extension of the function (this is always recommended!) and beks in mind the criterion just formulated, then the nature of the behavior of the Fourier series at the end points of the interval will be immediately apparent.

I I. Even and Odd Functions

Let the function f(x), defined either on the whole x-axis or on some interval, be symmetric with respect to the origin of coordinates. We say that f(x) is an even function if

for every x. This definition implies that the graph of any even function y = f(x) is symmetric with respect to the y-axis [see Fig. 12(a)]. It follows

4

Page 36: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

22 TRIoONOMglrUC FOURIER SERIES CHAP. 1

from the interpretation of the integral as an area that for even functions we have

for any I, provided that f(x) is defined and integrable on the interval [-I, I]. We say that the function f(x) is odd if

for every r In particular, for an odd function we have

f(-0) = -f(O).

so that f(0) = 0. The graph of any odd function y = f (x) is symmetricwith respect to the point 0 [see Fig. 12(b)]. For odd functions

for any I, provided that f(x) is defined and integrable on the interval [- 1, I]. The following properties are simple consequences of the definition of even

and odd functions : (a) The product of two even or odd functions is an even function; (b) The production of an even and an odd function is an odd function. In fact, if cp(x) and +(x) are even functions, then for f(x) = p(x)+(x), we

have

while if p(x) and +(x) are odd, we have

This proves Property (a). On the other hand, if cp(x) is even and +(x) is odd, then

which proves Property (b).

12. Cosine and Sine Series

Let f(x) be an even function defined on the interval [ - x , x ] , or else an even periodic function. Since cos nx (n = 0 , 1,2, . . .) is obviously an even function, then by Property (a) of Sec. 1 1 the function f (x) cos nx is also even. On the other hand, the function sin nx (n = 1,2, . . .) is odd, so that the function f (x) sin nx is also odd, by Property (b) of Sac. 11. Then, using

Page 37: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

(6.5), (1 1.1) and (1 1.2). we find that the Fourier coefficients of the even function f (x) are

Therefore, the Fourier series of an even function contains only cosines, i.e., 00

f(x) - + 2 an cos nx, n o 1

where the coefficients a,, are given by the formula (12.1). Now, let f(x) be an odd function, defined on the interval [-x, XI, or else

an odd periodic function. Since cos nx (n = 0, 1,2, . . . ) is an even function, the function f(x) cos nx is odd, by Property (b) of Sec. 1 1, and since sin nx (n = 1.2,. . .) is odd, the function f(x) sin nx is even, by Property (a) of Sec. 11. Then, using (6.5), (1 1.1), and (1 1.2), we find that the Fourier coefficients of the odd function f(x) are

Therefore, the Fourier series of an odd function contains only sines, i.e.,

f (x) - 2 bn sin nx, n=1

where the coefficients bn are given by the formula (12.2). Since the Fourier series of an odd function contains only sines, it obviously vanishes for x = -x, x = 0, and x = x (and in general for x = h), regardless of the values of f(x) at these points.

A problem which often arises is that of making an expansion in cosine series or sine series of an absolutely integrable function f(x) defined on the interval [0, x]. To expand f(x) in cosine series, we can reason as follows: Make the even extension of f(x) from the interval [0, X] onto the interval [ o x , 0] [see Fig. 13(a)]. Then all the previous considerations apply to the even extension off (x), so that its Fourier coefficients can be calculated by the formulas

Page 38: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

24 T R I O O N O ~ C POURIER SERIES CHAP. 1

which involve only the values of f(x) in the interval [0, x]. Therefoe, for computational purposes, there is no need to actually make the even extension off (x) from [0, n] onto [ - rr, 01.

To expand f(x) in she series, we first make the odd extension of f(x) from the interval [0, x] onto the interval [-z, O] [see Fig. 13(b)]. In doing so, the oddness requires that we set f (0) = 0. Then, the previous considera- tions again apply to the odd extension off (x), so that its Fourier coefficients are given by the formulas

a,,=O (n=O,1,2 ,... ), (1 2.4)

b,, = x - * ~ f ( x ) s i n n x d x o (n=1,2, ...),

which involve only the values of f(x) in the interval [0, n]. Therefore, as in the case of cosine series, there is no need to actually make the odd extension off (x) from [O, z ] onto [-z, 0 1.

However, in order to avoid mistakes in using the convergence criterion of Sec. 10, it is still recommended that a sketch be made of the function f(x) and its even (or odd) extension onto the interval [-z, 01, as well as of its periodic extension (with period 2x) onto the whole x-axis. This sketch will help in investigating the behavior of the "extended" function, which is the function to which the convergence criterion has to be applied.

13. Examples of Expansions in Fourier Series

Example 1. Expand f(x) = x2 (-x < x < x) in Fourier series. The function f(x) is even; the graph of f(x) together with its periodic extension is shown in Fig. 14. The extended function is continuous and piecewise smooth. Therefore, by the criterion of Sec. 10, its Fourier series converges to f(x) = x2 everywhere in [- rr, n], and converges to the periodic extension of f(x) outside [-z, x]. Moreover, the convergence is absolute and uniform.

Page 39: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

A calculation shows that

Furthermore, integrating by parts, we find that

4 x-lc = - [X ms nx] - 5 cos nx dx m2 X=O xn2 o

while b, = 0 (n = 1.2, . . . ), since f(x) is even. Therefore, for -x < x 4 rr, we have

x 2 , - - 4 cos x - cos 2x + cos 3x - . . .). 3 " ( 22 32

Example 2. Expmd f(x) = 1x1 ( - x < x < x ) in Fourier series. The hnction f(x) is even; Fig. 15 shows the graph of f(x) together with its periodic extension. The extended function is continuous and piecewise smooth, so

Page 40: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

26 TRIOONOMBTRXC FOURIER SERIES CHAP. 1

that the criterion of Sec. 10 is applicable. Therefore, its Fourier series converges to f (x) = 1x1 everywhere in [-x, n] and converges to the periodic extension of f(x) outside [-x, x]. Moreover, the convergence is absolute and uniform.

Since 1x1 = x for x 3 0, we have

an = ~ ~ X C ~ S I U C ~ X = x o -- 7 ~ 1 0 1. sin nx tix

L -- - -

L =- [COS nx] = - [cos m - 11 m2 %PO m2

It follows that a, = 0 for even n, and that an = -4Jxnz for odd n. Finally, b,, = 0 (n = 1,2, . . .), since f(x) is even. Thus, for - x ( x < x, we have

Example 3. Expand f(x) = rsin xl tr Fourier series. This function is defined for all x, and represents a continuous, piecewise smooth, even func- tion. Its graph is shown in Fig. 16. The criterion of Sac. 10 is applicable, and hence f(x) = lsin X I is everywhere equal to its Fourier series, which is absolutely and uniformly convergent.

Since lsin XI = sin x for 0 ( x < n, we have

4 a. = 21. sin x tix = 3 tt 0 R

and

Page 41: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

- - (" [sin (n + 1)x - sin (n - l)x] dr 7c 0

for n # 1, while for n = 1

Moreover, b, = 0 (n = 1,2, . . .), since f (x) is even. Therefore, for all x we have

2 4 cos2x cos4x cos 6x lsin xl = - - - ( + 15 + 35 + . . .).

x n 3

Example 4. Expand f(x) = x ( - x < x < X ) in Fourier series. The function f (x) is odd ; Fig. 17 shows the graph off (x) together with its periodic extension. The extended function is piccewise smooth and discontinuous at the points x = (2k + 1)tr (k = 0, f 1, f 2,. . .). The test of Sec. 10 is applicable, and the Fourier series of f(x) converges to zero at the points of discontinuity.

Since f (x) is odd

2 x=II = - -

nn [x cos nxl + 2 r cos nx tix x-0 7cn 0

Page 42: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

CHAP. 1

Therefore, for -IT < x < sc, we have

Example 5. Expand f(x) = 1 (0 < x < IT) in sine series. Making the odd extension of f (x ) onto the interval [-sc, O] produces a discontinuity at x = 0. Figure 18 shows the graph off (x) and its odd extension, together with its subsequent periodic extension (with period 2lc) over the whole x-axis. The convergence criterion of Sec. 10 is applicable to this "extended" function. Therefore, its Fourier series converges to f(x) = 1 for 0 < x < z. Outside the interval 0 < x < lc, it converges to the function shown in Fig. 18, with the sum of the series being equal to zero at the points x = h ( k = 0, f 1, * 2 ,... ).

Since

bn=2r sc o sin nx &c

we have

sin 3x sin Sx sinx+- + 5 3

for 0 < x < ~t.

Example 6. Expand f(x) = x (0 < x < 274 r) Fourier series. This example bears a superficial resemblance to Example 4, but the difference is immediately apparent if we construct the periodic extension of f(x) (see Fig. 19). The criterion of Sec. 10 is applicable to this extended function. At the points of discontinuity, the Fourier series converges to the arithmetic mean of the right-hand and left-hand limits, i.e., to the value R. The function f(x) is neither even nor odd.

Page 43: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Since

b, = I j2= x sin nx d~ TC 0

1 x=2= - - - - 2 1% cos nx] + L 12= cos nx dx = -9 7tll n n o n

we have

x = x - 2 sin 2x sin 3% 2 + 3

for 0 < x < 27~.

Example 7. Expand f(x) = xz (0 < x < 2%) in Fourier series. This example resembles Example 1, but the graph of the periodic extension of f(x) immediately shows the difference (see Fig. 20). The criterion of Sec. 10 is

Page 44: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

30 TRIGONOMETRIC FOURIER SERIES CHAP. 1

applicable, and at the points of discontinuity the series converges to the arithmetic mean of the right-hand and left-hand limits, i.e., to the value 2x2. The function f(x) is neither even nor odd.

Since

2 x= 2lr =- 4

[x cos n x ~ - 2 J2% cos nx du = 7 fvt2 X=O 1~12 o n2

b,, =1r z 0

x2 sin nx dx

1 x-2rr = - - xn [x2 cos nxl + 2 p x cos fix tix

XPO 7En 0

we have

~2 = - cosx - xsinx + cos 2x x sin 2x - 3

+... 22 2

cosnx xsin nx + n2

- n

sin nx ,

for 0 < x < 2%.

Example 8. Expand f(x) = Ax2 + BX + C (-n < x < n), where A, B, and Care cowtmrts, in Fmrier series. The graph of f ( x ) is a parabola. By periodic exteasion, we can obtain a continuous or a discontinuous function, depending on the choice of the constants A, B, and C. Figure 21 shows a possible extension for cortain values of A, B, and C.

We could calculate the Fourier coefficients from the appropriate formulas, but there is no need to do so, since we can use the expansions for the functions x2 and x (-IT < x < n), given in Examples 1 and 4. The result is

An2 PO cos nx A ~ ~ + B X + C = - + C + ~ A Z ( - I ) " - - 3 2B 2 (- 1)" sin nx .

n o 1 n2 n o 1 n

Page 45: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 13 TRIGONOMETRIC FOURIER SERIES 3 1

Example 9. Expand f(x) = Ax2 + BX + C (0 < x < 2%) in Fourier series. Figure 22 shows the periodic extension off (x) for a certain choice of the constants A, B, and C. Using the expansions of the functions x2 and x (0 < x c Zz), given in Examples 6 and 7, we find that

QO sin nx - (4zA - 2B) 2 - n= 1 n

for 0 < x < 2%.

We can use these examples to calculate the sums of some important trigonometric series. For example, (13.5) immediately gives

and from (13.5) and (13.6). we infer that

cos nx 3x2 - dnx + 2n2 t n2 - -

12 . ( O < x < 2 x ) . (1 3.8) n = l

Since the terms of the series on the left do not exceed lln2 in absolute value, the series is uniformly convergent, which means that its sum is continuous

Page 46: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

for all x (see Sec. 4). Therefore, (13.8) is valid for 0 < x < k, and not just for0 e x < % .

Similarly, (1 3.3) gives

sin nx x '2 ( - x < X < n),

n o 1 n

(13.1) gives

cosnx n2 - 3x2 12 ( o n < x < n), (1 3.10)

and (13.2) gives

8 (0 c x ( x). (1 3.12)

Moreover, subtracting (1 3.1 1) from (1 3.7). we obtain

- - "-" 4 (0 < x < n),

n= 1

and subtracting (13.2) from (13.8), we obtain

These formulas also allow us to calculate the sums of some numerical series, For example, if we set x = 0, (13.8) and (13.10) become

while if we set x = z/2, (1 3.1 1) becomes

14. The Complex Form of a Fourier Series

Let the function f(x) be integrable on the interval [on, n], and form its Fourier series

Page 47: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 14 TRIGONOMETRIC FOURIER SERIES 33

00

f(x) 9 + 2 (an cos nx + bn sin nx), n= 1

We shall make use of Euler's well-known formula, relating the trigonometric and exponential functions :

elf@ = cos cp + i sin cp.

It is an immediate consequence of this formula that

&w + e-iv eicp - e-icp cos cp = 9 sin cp = 2

. 2i

Therefore, we can write einx + e-i8.x

cos nx = 2 9

el= - r i m - eim + e - fn~ sin nx = = i .

2i 2

Substituting these expressions in (14.1). we obtain

If we set

then the mth partial sum of the series (14.3), and hence of the series (14.1). can be written in the form

Therefore it is natural to write

This is the complex form of the Fourier series of f(x). The convergence of the series (14.6) must be understood to mean the existence of the limit as m -t ao of the symmetric sums (14.5).

Page 48: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

34 TRIGONOMETRIC FOURIER SERIES CHAP. 1

The coefficients c,, given by (14.4) are called the complex Fourier co- eficients of the function f(x). They satisfy the relations

In fact, by Euler's formula and (14.4). we have

1 1 dA f(x)r'u dx = - 27t [ r - A fix) cos nr dx - i [f(x) sin nr h]

= +(a,, - ibJ = cn

for positive indices and

1 1 5 f(x)* = - 2 n [ K f (x) cos nx d* + i r - f t f (x ) sin nx dx]

for negative indices. It is useful to bear in mind that iff (x) is red, then the coefficients c,, and c-,, are complex conjugates. This is an immediate consequence of (1 4.4).

Incidentally, we note that the formula (14.7) can also be obtained directly, just as the formulas (14.2) were (see Sec. 6), if we assume that the sign = appears in (14.6) instead of the sign - and that term by term integration is legitimate. In fact, multiplying both sides of the equality

by e" and integrating term by term over the interval [-rr, x], we obtain

since for k # n (see Sec. 5) we have

IC

= 4 [OOS (k - n)x + t sin (k - n)x] dx = 0,

is., all the integrals on the right vanish except the one corresponding to the index k = n, while for k = n, we obtain the number 2%. The fonnula (14.7) is an immediate consequence of (14.8).

Page 49: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

TRIGONOMETRIC FOURIER SERIES 35

IS. Fynctions of Period 21

If it is required to expand a function f(x) of period 2l in Fourier series, we set x = lt/n, thereby obtaining the function cp(t) = f(lt/n) of period 2% (see Sec. 3). For cp(t) we can form the Fourier series

00

~ ( t ) .Y + 2 ((I. cos nt + bn sin nt), n = l

1 = bn = ; /-= ~ ( l ) sin nt dt = jZn f e) sin nt (n = 1.2,. . .).

Returning to the original variable x by setting t = rrxll, we obtain

m x n= 1 I f(x) - q + 2 (an cos - + bn sin -

where

b,= lr f(x) sin mx - dx I -1 I

The coefficients (15.3) are still called the Fowier coeficients of f(x), and the series (15.2) is still called the Fourier series of f(x). If the equality holds in (15.1), then the equality holds in (15.2), and conversely.

We could have constructed a theory of series of the form (15.2) directly, by starting from a trigonometric system of the form

7cx XX m x m x 1, cos -9 sin -9.. ., cos - I 1 9 sin - 1 '"" (1 5.4)

just as we did in the case of the basic trigonometric system (5.1). The system (15.4) consists of functions with the common period 21, and it is easily verified that these functions are orthogonal on every interval of length 21. The considerations of Secs. 6,7, 10, 12, and 14 can be repeated as applied to the system (15.4), and the result is a formulation analogous to that given in these sections, except that x is replaced by I. In particular, instead of a function f(x) of period 21, we can consider a function defined only on the interval [-I, I] [or on any other interval of length 21, provided we appro- priately change the limits of integration in (1 5.3)]. The Fourier series of such a function is identical with that of its periodic extension onto the whole

Page 50: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

36 TRIGONOMETRIC FOURIER SERIES CHAP. 1

x-axis. The convergence criterion of Sec. 10 continues to "work," if we replace the period 2n by the period 2l.

If f(x) is even, the formulas (15.3) become

while if f(x) is odd, they become

As in Sec. 12, we can use this fakt to expand a function f(x) defined only on the interval [0, I] in cosine series or in sine series (making the even or the odd extension off (x) onto the interval [- I, OD.

The complex form of the series (15.2) is

where

a0 a,, - ibn - 9 C-n - an + ibn

co = T' = 2 2 (?I = I, 2,. . .).

Example 1. Expand the fMcrion f (x), &fined by

for 0 6 x < 1

f(x) = 2'

I 6 for p x < 1

in cosine series. Figure 23 shows the graph of f(x) and its even extension onto the interval [ - I, 01, together with its subsequent periodic extension

Page 51: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 1s TRIGONOMETRIC FOURIER SERIES 37

(with period 22) onto the whole x-axis. The convergence criterion can obviously be applied everywhere.

For 212 < x g 2, we have f(x) = 0, so that

Making the substitution = t, we obtain

a, = - 1 12 r I 2 ws t ms nt dt = - r [cos (n + 111 + cos (n - l)tl dt,

7t 0 X 0

whence

(n + l)t + sin (n - l)t] t=nfl an = - (n > 1). n - 1 t-0

Therefore, for odd n > 1

a, = 0, while, for even n

Thus we have

I for 0 < x < ~

1 1 7tx 2 " (-1)" -+zcos- - - 2 4n2 - 1 COS- = x X n = ~ I

for - < x 6 I. 2

This series converges on the whole x-axis to the function shown in Fig. 23.

Example 2. Expand the function f(x), defned by

2 for 0 ( x b z ~

I 1 - x for z < x < l ,

in sine series. Figure 24 shows the graph of f(x) and its odd extension onto the interval [-I,O], together with its subsequent periodic extension (with

Page 52: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

38 TRIGONOMETRIC FOURIER SERIES CHAP. 1

period 21) onto the whole x-axis. The convergence criterion can be applied everywhere.

In this case, we have

an=O (n=O, l ,2 ,... ),

Setting xx/ l= t, we obtain

bn = 1"" t sin nt dt + 2 J1 (n - t ) sin nt dt 7 9 0 x2 x/2

cos nt dt t =O

+ 21 [ - (1. - t )msnt l tnn - EJ= l t 2 n cos nt dt

2 n2n 4 2

41 =- 7en sin -= x2n2 2

Therefore

I for 0 4 x G r 3nx I + - s i n -. . .

1 32 I 52 I I 1 - x for z < x ( l .

This series converges on the whole x-axis to the function shown in Fig. 24.

PROBLEMS 1. Expand the following functions in Fourier series:

a) f(x) = (-rc < x < z), wherea # Oisaconstant; b) f(x) = cos ax ( - x 4 x 4 z), where a is not an integer;

Page 53: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

PROBLEMS TRIGONOMETRIC FOURIER SERIES 39

C) f(x) = sin ax (- x < x < x), where a is not an integer; 0 for - 7 c < x < o ,

= {x for o < x < 7c.

2. Using the expansion of Prob. 1 b, show that

1 3 -

sins z

where z is any number which is not a multiple of rr.

3. Using the expansion of Prob. la, expand the following functions in Fourier Series:

a) The hyperbolic cosine @x + e-ax

cosh x = 2 (-n < x < rr);

b) The hyperbolic sine @X - e-4X

sinh x = 2 (0% < x < x).

4. Expand the following functions in Fourier cosine series:

a) Ax) = sin ax (0 4 x g rr), where a is not an integex; 1 for 0 Q x Q h,

b) fb) = {o f or h < x < x ; 4

c) f(x) = I - 3% for 0 Q x 4 2h,

10 for 2h < x < x.

5. Expand the following functions in Fourier sine series:

XX I sinl for O < x < T I

for - < x < 1; 2 rrx 1 sinT for O < x < r

1 for < x < 1.

6. Expand the periodic function

in Fourier series.

7. Let f(x) haw period 2% and let 1 f(x) - f@)l < clx -.yla, for some constants

Page 54: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

40 TRIaONOMBTRIC FOURIER SERIES CHAP. 1

c > 0, a > 0, and for all x and y. f(x) obeys this inequality, it is said to satisfy a H'l&r (or Lipschitz) condition of order a,] Show that

where an and b, are the Fouriq aeiiicients of f(x).

8. Expand the following functions in Fourier sine series: a) f ( x ) = c o s x ( O < x < r r ) ; b) f(x) = x3 (0 < x < rr).

9. Let f(x) be a function of period 2% defined for - x < x < x. Let f (x) have the Fourier series

$ + 5 (anmrnx + bnsionr), n = l

and let

Show that fdx) is an even function and A(*) an odd function, with Fourier SeTies

a0 a0

respectively, Show that the function f(x - .rr) has the Fourier series 00

3 + 2 ( - I)n(a, cos nx + bn sin nx). 2 n o 1

10. Sum the Series

by using Example 8 of Sec. 13 and the results of the preceding problem.

11. Find the sum of each of the following numerical series by evaluating at a suitable point a Fourier series given in the text or in the problems:

"O sin nh

n=1 n = l

12. Show that the Fourier series for the function f(x) = x on the interval -R < x < x docs not converge uniformly, but that the Fourier series for the function f(x) = x2 does converge uniformly. Find the Fourier series for f(x) = x4 by integrating the Fourier series for f(x) F x2 between the limits 0 and x.

Page 55: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

ORTHOGONAL SYSTEMS

I. Definitions

An infinite system of real functions

fPo(x), p1(x), fPz(x), . , cpXx),

is said to be orthogonal on the interval [a, b] if

We shall always assume also that

The condition (1.2) says that every pair of functions of the system (1.1) is orthogonal (see Ch. 1, Sec. 5), while the condition (1.3) says that none of the functions of the system is identically zero.

We have already encountered special cases of orthogonal systems, i.e., the basic trigonometric system

1, cos x, sin x, . . ., cos nx, sin nx, . . . , (1 -4) which is orthogonal on any interval of length 2n, and the more general trigonometric system

'IUC 7tx mcx nnx 1, cos - sin -, . . . , cos 7, sin - 1 ' I 1 ' * * "

which is orthogonal on any interval of length 2l (see Ch. 1, Secs. 5 and 15). 41

Page 56: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

CHAP. 2

The system (1.1) is said to be normalized if

Evary orthogonal system can be normalized, i.e., one can always choose the constants h, b,p,. . ., pn,. . . in such a way that the system

which is obviously still orthogonal, is now also normalized. Such a system is said to be orthonormal. In fact, the condition

implies that

1

Finally, we introduce the notation

and call this number the norm of the function vn(x). If the system (1.1) is normalized, then obviously

Ilpnll=l (n-0,1,2, ...).

2. Fourier Series with Respect to an Orthogonal System

We now essentially repeat the considerations of Ch. 1, Sec. 6, in a more general context. Let f(x) be a function defined on the interval [a, b]. Sup- pose that f(x) can be represented as the sum of a series involving the functions of the orthogonal system (1.1), i.e., suppose that everywhere on [a, b]

where co, cl, . . ., c , . . . are constants. To calculate these constants, we assume that the series

obtained by multiplying the equations (2.1) by vn(x), can be integrated term

Page 57: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 2 ORTHOGONAL SYSTEMS 43

by term over the interval [a, b]. According to (1.2), this integration gives

Therefore we have

Now suppose that we are given a function f(x) defined on the interval [a, b], and we wish to make a series expansion of f(x) with respect to the functions of the system (1. l), without knowing in advance whether or not such an expansion is possible. If such an expansion is actually possible (and if the required term by term integration is also possible), then, as we have just seen, we must arrive at the formulas (2.3). Therefore, in trying to find the required expansion of the function f(x), it is natural to first examine the series with coefficients given by (2.3), and see whether this series might turn out to converge to f(x). The coefficients given by (2.3) are called the Fourier coeficients off@) with respect to the system (1. I), and the correspond- ing series is called the Fourier series of.f(x) with respect to the system (1.1). If the system (1.1) is normalized, then the formulas for the Fourier coefficients take a particularly simple form:

Until it has been ascertained that the Fourier series of f(x) actually converges to f(x), we write

However, it should be noted that even in the case where the Fourier series turns out to be divergent (and this actually happens sometimes), the Fourier series still has various remarkable properties which will be discussed below.

If the functions of the system (1.1) are continuous, and if the series in the right-hand side of (2.1) is uniformly convergent, then it is easy to prove that the series (2.2) is also uniformly convergent and can therefore be integrated term by term (see the proof of Theorem 1 of Ch. 1, Sec. 6). This immediately implies the following

THEOREM. If the fknctions of the system (1.1) are continuous and if the series expansion (2.1) off (x) is uwrmly convergent, thm (2.1) is the Fourier series of f(x).

Page 58: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

$4 ORTHOGONAL SYSTEMS CHAP. 2

3. Some Simple Orthogonal Systems

In addition to the orthogonal systems (1.4) and (1.5) just cited, we consider the following systems:

I. The system

1, COS X, COS 2x, . . . , cos nx, . . . is orthogonal on the interval [0, x]. In fact

which means that the functions cos nx and 1 are orthogonal. Moreover, we have

Koosnxcwmxclx = 1 [cos (n + m)x + cos (n - m)x] dx

which follows from (3.1). This proves that the system I is orthogonal. In writing Fourier series with respect to the system I, we continue to use

the notation introduced in Ch. 1, i.e., we write

+ alcosx + a2cos2* + * * a + ~ , , C O S ~ X + * * a . f(x) - q With this notation for the Fourier coefficients, the formulae (2.3) give

and

Since

we can write

a,, = - 2rf(x)cosnxdx (n=0,1, 2,...). = 0

Page 59: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 3 ORTHOOONAL SYSTEMS 45

Thus, as is to be expected, we arrive at the same formulas (12.3) as obtained in Ch. 1 for cosifie series.

11. The system

sin x, sin 2x, . . . , sin nx, . . . is orthogonal on [0, n]. In fact

for n # m [see (3.1)]. As in the preceding case, in writing Fourier ~eries with respect to the system 11, we continue to use the notation adopted in Ch. 1 :

f(x) bl sinx + b2 sin 2x +- + bnsin nx + Then by (2.3)

sin nx dr b,, = (n = 1,2,. . .). I]I sin2 nx dx

Moreover, since

we have

i.e., as is to be expected, we arrive at the formulas (12.4) obtained in Ch. 1 for sine series.

111. The system

sin x, sin 3x, sin Sx, . . ., sin (2n + l)x, . . . is orthogonal on [O, ~121. In fact, for n # m and n, m = 0, 1,2, . . . , we have

c2 sin (2n + 1)x sin (Zm + 1)x dx

= 1 [COS 2(n - m)x - cos 2(n + m + I)X] dx 2 0

=- [ xu@ 1 sin 2(n + m + 1)x *nI2 [ ] -0. sin 2(n - m)X] x=o - i 2(n + m + 1) ,o 2 2(ri - m)

The formulas (2.3) fdr the Fourier coefficients give

Page 60: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

CHAP. 2

JY2 f (x) sin (20 + 1)x dr Cn = (n = 0, 1,2,. . .).

K I 2 sin2 (2n + 1)x dx

But

1'' sin2 (b + I)X = J =/z 1 - cos (4n + 2)x dr x 0 2 = 3'

and therefore

We now show that if f(x) is a function defined on the interval [0,7r/2], then we can amve at the expansion of f(x) with respect to the system I11 by starting from the basic trigonometric system, just as in Ch. 1, Sec. 12, we amved at the expansions in cosine series or sine series of a function defined on [0, x] by using its even or odd extensions onto the interval [-x, 01. To do this, we have to generalize the concepts of evenness and oddness of a function. Thus, let f(x) be defined either on the whole x-axis or on some interval which is symmetric with respect to the point x = 1. We shall say that f(x) is even with respect to x = 1 if

for every h. This means that the curve y = f(x) is symmetric with respect to the line x = 1 (see Fig. 25).

If a function f (x ) is even with respect to x = I, then obviously

f (x) dk = 2 1' f(x) dx, I-a I-a

and in particular (for a = 1)

Page 61: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC, 3 ORTHOGONAL SYSTEMS 47

Similarly, we say that f(x) is odd with respect to x = 1 if

for every h. This means that the curve y = f(x) is symmetric with respect to the point (1,O) (see Fig. 26). If a function f(x) is odd with respect to x = I, then obviously

sl+' f (x) dx = 0, I-a

and in particular

With this interpretation of evenness and oddness, we can assert that the product of two even or two odd functions is even, whereas the product of an even function and an odd function is odd. The proofs are essentially the same as those given in Ch. 1, Sec. 1 1.

Now let f(x) be defined on the interval [O,7c/2], and make the even extension off (x) onto the interval [x/2, IT] (see Fig. 27). This gives a function

g(x) defined on [0, x ] which coincides with f(x) on [0, x/2]. Expand the function g(x) in a cosine series, which is equivalent to making the odd extension of g(x) onto the interval [-x, 01 (see Ch. 1, Sec. 12) ; the result is

Page 62: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

48 O I I ' W O C S O N A L S Y ~ CHAP. 2

Next, we observe that the functions of the system III are even with respect to x = 4 2 . In fact, for n = 0,1,2,. . ., we have

sin (t, + 1) 6 - h)

= sin (2, + l)z=cos (2n + 1)h

- cos (Zn + 1) iosin (2n + 1)h

+ cos (2n + 1) f asin (2n + l)h = sin (2n + 1) 6 + h),

since

Therefore, since the functions g(x) and sin (2n It- 1)x are even with respect to x = ~12, it follows from (3.3) and (3.4) that

bWt = 2 r &) sin (2, + 1)x dw X 0

=:y % 0

f (x) sin (2n + 1)x dw

On the other hand, the functions sin 2nx (n = 1,2, . . .) are odd with respect to x = z/2, since

sin 2n 6 - h) = sinm cos 2nh - cosm sin 2nh

= - (sin m cos 2nh + coa xn sin 2nh)

Therefore, the product g(x) sin Znx (n = 1,2, . . .) is odd with respect to x - x/2, and hence

Thus, finally, we have expanded the function g ( ~ ) , and consequently the firnction f(x), in a sine series such that all the coefficients with even indices

Page 63: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 3 ORTHOGONAL SYSTEMS 49

vanish. The coefficients with odd indices are given by the formulas (3.4), which coincide with (3.2).

We have gone through this rather lengthy argument in order to be able to apply the convergence criterion of Ch. 1, Sec. 10 (which was formulated for functions of period 2%) to the case of series expansions with respect to the system 111. Our argument shows that this convergence criterion must be applied to the function obtained from f(x) by first making an even extension of f(x) onto [7t/2, rr], then making an odd extension of the -resulting function onto 1-5 01.1 and finally extending the result periodically (with period 2%) onto the whole real axis.

IV, V. The system

7tX 27cx mrx 1,cos- cos-, ..., COS- I ' I 1 ' * "

and 7tx %x n7tx sin-, sin -, . . . , sin - I I 1 " * *

are orthogonal on the interval [0, I]. In fact, if we make the substitution zx/ l= f, the integrals of pairs of functions from each of these systems redua to the corresponding integrals for the systems I and 11.

VI. The system

nx . 37tx Sscx 9 sin - sin -9 sin - sin (2n + 1)xx

2l 21 " * " 22 ,*..

is orthogonal on the interval [0, I]. In fact, making the substitution 7tx/21 = t, we obtain

(2n + 1)xx (2m + l ) m h J' sin 21

sin 0 21

since everything has been reduced to the orthogonality of functions of the system 111. For the Fourier coefficients with respect to the system VI, we obtain

(2n f J)m dx Jif(x) sin 2

c,, = = 2 1' f(x) sin (h + l)ruCd*. Ji sin2 (a +* J)ILx dX l o 21 21

If we want to expand a fundion f(x) defined on [0, I] in a Fourier series

1 Figure 27 shows the graph of y = f (x) together with the indicated twofold extension.

Page 64: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

50 ORTHOCSONAL SYSTEMS CHAP. 2

with respect to the system VI, we first make the substitution zx/W = t. This reduces the problem to expanding the function ~ ( t ) = f(Z*), defined on [O,n/2], with respect to the system 111. Then, we return to the required expansion with respect to the system VI by transforming back to the variable x. It follows that we can apply the convergence criterion of Ch. 1, Sec. 10 to expansions with respect to the system VI (which are quite often encountered in the applications), since the criterion is applicable to the system 111.

Later, we shall deal with orthogonal systems consisting of functions which are more complicated than trigonometric functions (e.g., Bessel functions).

4. Square Integrable Functions. The Schwan Inequality

We shall say that the function f(x) defined on the interval [a, b] is squme integrable if f(x) and its square are integrable on [a, b]. Every bounded integrable function must be square integrable, but this is not always the case for unbounded integrable functions. For example, the integral

exists, but the integral

does not exist. Let cp(x) and (~(x) be square integrable functions defined on [a, b]. First

we note that it follows from the elementary inequality

l d l +b2 + 4J2)

that the function I&( is integrable? Now consider the inequality

which holds for an arbitrary constant A, and set

Then

2 Incidentally, this implies that a square integrable function is always absolutely integrable. (As we have seen, the converse is not always true.) To see this, it is sufEcient to set +(x) = 1.

Page 65: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

for any A. Therefore, the graph of the polynomial

p = A + Z B X + C X 2

is a parabola lying above the x-axis or perhaps touching it (see Fig. 28).

It follows that our polynomial cannot have distinct real zeros, since then the curve would intersect the x-axis in two points. Therefore, the dis- criminant of the polynomial must satisfy the inequality

B2 - AC < 0,

B2 < AC.

Recalling the meaning of A, B, and C, we obtain

This very useful inequality is called the Schwarz inequality.3 Using the Schwan inequality, it is easy to show that the sum of a finite

number of square integrable functions is also a square integrable function. In fact, for two functions we have

and the generalization to the sum of any number of functions is straight- forward.

5. The Mean Square Error and i t s Minimum

Let f (x) be an arbitrary square integrable function defined on the interval [a, b], and let an@) be a linear combination of the first n + 1 functions of the system (1. I), i.e.,

an@) = Yoc~dx) + YICPI(~) + + Yn~n(x), (5.1)

3 Called the Bunlakmsk1 ineqrsalitv by Russian mathematicians. (Trunslotor)

Page 66: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

52 ORTHOGONAL S Y ~ M S CHAP. 2

where yo, yl,. . ., yn are constants. By hypothesis, all the functions vn(x) of the system (1.1) are square integrable [see (1.311. Therefore, the linear combination an(#) and the difference f(x) - an(x) [n = 0, 1,2, . . .] are also square integrable functions.

Now consider the quantity

which we call the mean square error in approximatingf(x) by an(x). There are rdany ways of calculating the deviation of the linear combination o,(x) from the function f(x); we have chosen the mean square error, which is particularly appropriate in the theory of Fourier series.

We now pose the problem of choosing the coefficients yo, yi, . . . , yn (for a given n) in such a way that the mean square error 8, is a minimum. It follows from (5.2) that

8, = lb f qx) dx - 2 lb f(x)a,,(x) tak + Jb d,(x) dx. u a a

By (5.1) we have

But according to (2.3)

where the ck are the Fourier coefficients of the function f(x), and therefore

Furthermore, we have

Tbe last sum extends over all possible unequal indice8 p and q that do not exceed n. By the orthogonality of the system (1. l), this sum vanishes. Thus, we find that

Page 67: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Substituting (5.4) and (5.5) in (5.3), we obtain

Here

i.e., do not depend on yo, yl, . . . , y , and therefore the quantity 6, is obviously a minimum when

which is equivalent to the conditions

y k = c k (k=O,l ,2 ,..., n).

Thus, finally, the mean square error is a minimum when the coeflcients in the linear combination (5.1) are the Fourier coeflcients. Denoting the minimum value of the mean square error by A,,, we have

This expression shows that as n increases, the nonnegative quantity A,, can only decrease. Thus, as n increases, the partial sums of the Fourier series give a closer approximation to the function f(x), i.e., an &pproximation with smaller mean square error.

6. Bessel's Inequality

Since A,, 3 0, it follows from (5.6) that

where n is arbitrary. The sum on the right can only increase as n increases.

Page 68: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

54 ORTHOOONAL SYSTEMS CHAP. 2

Therefore, since it is bounded by a constant (the integral on the left), the sum has a finite limit as n + oo. Thus, the series

converges and

This very important result is called Bessel's inequality. It follows at once from the convergence of the series on the right that

lim k,Jyal = 0. wao

If the system (1.1) is normalized, then hssel's inequality takes the form

and therefore, the sum of the squares of the Fourier coefficients is convergent. For a normalized system, (6.2) becomes

lim c,, = 0, I)--+-

i.e., the Fourier coefficients approach zero as n + 00.

7. Complete Systems. Convergence in the Mean

The system (1.1) is said to be complete if for any square integrable function f(x), the equalitv

holds (instead of Bessel's inequulity). Here, as above, the ck (k = 0, 1,2, . . .) are the Fourier coefficients of the function f(x). The equality (7.1) is called the completeness condition for the system (1.1). The following simple result is an immediate consequence of the completeness condition:

THEOREM 1. k t f (x) and F(x) be square integrable functions for which

and let the system (1.1) be complete. Then we have

Page 69: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 7 ORTHOOONAL SYSTEMS 55

Proof. The sum f(x) + F(x) and the difference f(x) - F(x) are square integrable functions. Moreover, the sum has Fourier coeffi- cients ck + Ck and the difference has Fourier coefficients ck - C,. By the completeness relation, we have

and then subtraction gives

The following result has important consequences :

THEOREM 2. A necessary and suficient condition for the system (1 .I) to be complete 2r that the relation

hold for any square integrable fwt ion f (x), where the ck (k = 0, 1.2, . . . ) are the Fourier coeflcients o f f x ) with respect to the system (1.1).

Proof. Use the relation (5.6) and the fact that the completeness condition is equivalent to

If the relation (7.3) is satisfied, we say that the Fourier series converges to f(x) in the mean. Therefore, Theorem 2 can also be formulated as follows :

A necessary and suficient condition for the system (1.1) to be complete Is that the Fowier series of any square integrable function f(x) converge to f(x) in the mean.

It should be noted that ordinary convergence of a Fourier series to the function from which it is formed does not always occur, even if the system (1.1) is complete. Nevertheless, as we have just shown, convergence in the mean always occurs for complete systems (it is understood that we are talking about square integrable functions). In particular, these remarks apply to the trigonometric system (whose completeness will be proved in Ch. 5, Sec. 2).

Page 70: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

56 ORTHOOONAL SYSTEMS CHAP. 2

These remarks show the importance of the concept of convergence in the mean and suggest that this kind of convergence be regarded as a generalizatwn of ordinary convergence. To completely justify this approach, we now show that a Fourier series can converge in the mean to only one firnction (with a certain stipulation). Thus suppose that in addition to (7.3), the relation

also holds. Then, using the elementary inequality

we find that

By (7.3) and (7.4), this implies

and since the integrand is positive, it follows that

at the points of continuity of the integrand. But the integrand has only a finite number of points of discontinuity. Hence, the functions F(x) and f(x) coincide everywhere, except possibly at a 5 i t e number of points. Two such functions should hardly be considered different in the theory of Fourier series, since the values of a function at individual points can have no influence at all on the behavior of its Fourier series (since its Fourier coefficients are expressed in terms of integrals, and integrals do not depend on the values of the integrand at a finite number of points!). This allows us to draw the following conclusion :

THW)REM 3. If the system (1.1) is complete, then every square inte- grable function f(x) is completely determined (except for its values at a finite number of points) by its Fourier series, whether or not this series converges.

This means that no other function which is "essentially" different from

Page 71: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

a given function f(x), i.e., differs from f(x) at more than a finite number of points, can have the same Fourier series as f(x).*

8. Important Properties of Complete Systems

We now establish some very important properties of complete systems:

THEOREM 1. Jf the system (1.1) h complete, then any continuous fiction f(x) which is orthogonal to all the fiutctions of the system must be i&nticaUy zero.

ProoJ If f(x) is orthogonal to all the functions of the system, then all the Fourier coefEcients of f(x) vanish. Then the completeness relation (7.1) implies that

so that

f(x) = 0,

since f (x) is continuous.

THEOREM 2 if the system (1.1) is complete, if the functions of the system are conrinuous, and if the Fourier series of the continuous function f(x) is uniformly convergent, then the sum of the series equals f(x).

Proof. Let

and set

Since the functions of the system (1.1) are continuous, and since the series (8.1) is uniformly convergent, the sum of the series is continuous. It follows from the theorem of Sec. 2 that (8.1) is the Fourier series of s(x). Therefore, the continuous functions f(x) and s(x) have the same Fourier series. But then Theorem 3 of Sec. 7 implies that

f(x) = J(x), and by (8.1)

J It should be kept in mind that in this book integrable functions, and hence square integrable functions, are always assumed to be continuous except possibly at a finite number of points (Ch. 1, Sec. 4). Otherwise Theorem 3 would not be true. (Trmslator)

Page 72: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

58 ORTHOGONALSYSIBMS CHAP. 2

THEOREM 3. If the system (1.1) Ls complete, then the Fourier series of every square integrable function f(x) can be integrated term by term, whether or not the series converges.

In other words, if

f (4 - coTo(x) + ~ 1 9 1 ( ~ ) + ' + QPn(x) + , then

where xi and x2 are any points of the interval [a, b]. Proof. Assuming for definiteness that xl < XZ, we have

where we have used the Schwan inequality (see Sec. 4). By Theorem 2 of Sec. 7, the last term in (8.3) approaches zero as n + ao. Therefore,

which is equivalent to (8.2).

9. A Criterion for the Completeness of a System

In view of the importance of the concept of complete systems, it is appropriate to give a simple test for the completeness of a system. The following completeness criterion is very convenient:

If for every continuous Junction F(x) on [a, b] and any number r > 0, there exists a linear combination

on(x) Y ~ X ) + Y I ( P I ( X ) + + Y ~ X ) (9.1) Jbr which

then the system (1.1) Is complete.

Page 73: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 9 ORTHOOONAL SYSI'EUS 59

We begin by noting that given any square integrable function f(x), there exists a continuous function F(x) for which

This fact is quite clear geometrically, but for the reader who does not regard it as completely obvious, we give the following proof:

The function f (x) can have only a finite number of points of discontinuity. In particular, f(x) can have only a finite number of points at which it becomes unbounded. Every such point can be included in an interval of such small length that the sum of the integrals of the function f a(x) over these intervals does not exceed 44. Define the auxiliary function @(x) as being equal to f(x) outside these intervals and equal to zero inside them. @(x) is bounded and can have only a finite number of discontinuities, and obviously

Next, we include each point of discontinuity of Qyx) in an interval of such small' length that the total length I of all these new intervals satisfies the condition

where M is any number such that l@(x)l < M for a ( x $ b. Finally, consider the continuous function F(x) which equals @(x) outside

the intervals just described and which is "linear" inside each of them (see Fig. 29). Obviously we have

I l l - h i n t of

discontinuity

Page 74: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

60 ORTHOOONAL SYSTEMS CHAP. 2

It follows from (9.3). (9.4) and the inequality

(a + b)2 < 2(a2 + b9 that

< 2 Sf L((x) - @(x)l2 dx + 2 f [@(x) - F(x)J2 tk < s

i.e., the function F(x) satisfies the condition (9.2).

Returning to the proof of the completeness criterion, consider the linear combination o,,(x) which satisfies the inequality (9.1). Applying the in- equality (9.5). we obtain

where F(x) is the continuous function appearing in (9.2). We now recall that the linear combination whose coefficients are the Fourier coefficients c, gives the least mean square error (see Sec. 5). Therefore (9.6) gives

whence by (5.6)

which means that

as well. Since c is arbitrary, this gives (7. I), i.e., the system (1.1) is complete as asserted.

* 10. The Vector Analogy

Let i, j, and k be three space vectors of arbitrary length which are per- pendicular (orthogonal) to each other. If we want to expand a given vector r as a sum of the form

r = ai + bj + ck, (1 0.1)

Page 75: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

then the scalar coefficients a, b, c are calculated as follows: Take the scalar product of both sides of the equation (10.1) k t with i, then with j, and finally with k. Then, since these vectors am orthogonal, we obtain

(r, i) = alil2, (r, j) = bljlz, (r, k) = cIkI2,

so that

(We use the absolute value sign to denote the length of a vector.) Suppose now that we know the expansion (10.1) and we want to calculate

the length of the vector r. To do this, we take the scalar product of (10.1) with r. The result is

Id2 = 4r, i) + b(r, j) + ~ ( r , k), or, if we use the formula (10.2)

The quantities ali(, bljl, clkl are the projections of the vector r on the direc- tions of the vectors i, j, k. Therefore, the equation (10.3) is the familiar relation between the square of the length of a vector and its projections on three orthogonal directions. If besides the vector r, we consider the vector

and take the scalar product of r and R, then (10.1) and (10.4) give

(r, R) = aAIiI2.+ bBljlz + cCIk(2. (10.5)

The reader who has familiarized himself with the contents of this chapter will now immediately perceive an analogy between these vector considerations and our treatment of Fourier series. In fact, let us regard every square integrable function defined on the interval [a, b] as a generalized vector, and define the scalar product of two such generalized vectors by the formula

In particular, we have

Then, we regard the orthogonal system

Page 76: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

62 ORTHOGONAL SYSTEMS CHAP. 2

as a system of orthogonal vectors, in complete agreement with our definition of the scalar product. If we are given a square integrable functionf(x) and we want to represent f(x) as a series with respect to the system (10.6), i.e.,

then the argument which led to the relations (10.2) leads to the relations

which we recognize as the formulas for the Fourier coefficients [see (2.3)]. In fact, we have gone through precisely this derivation in Sec. 2. The reader will recognize that the completeness condition (7.1) is a generalization of the formula (10.3), and that (7.2) is a generalization of (10.5).

We now make some remarks concerning the term "completeness." Since my three-dimensional vector r can be represented in the form (10.1), i.e., as a linear combination of the vectors i, j, k, it is natural to call the system consisting of these three vectors complete. However, the situation is different if we try to represent an arbitrary three-dimensional vector r as a linear combination not of three orthogonal vectors but of just two vectors, say i and j Then, in general, we cannot write an equation of the form

r = ai + bj,

but the coefficients obtained from the formula (10.2) obviously satisfy the inequality

where the equality holds only if the vector r lies in the plane of the vectors i and j. Thus, two orthogonal vectors are not sujjfcient to be used in this way to represent any space vector, and therefore we say that a system of two orthogonal vectors is fncomplete.

Similar considerations apply to the case of expansions in Fourier series. If the system (10.6) satisfies the condition (7.1). then the system is "rich enough in functions" to permit every square integrable function to be represented by its Fourier series (in the sense of mean convergence). In this case, we say that (10.6) is a "complete system." However, if the system (10.6) does not satisfy the condition (7.1). then we say that it is incomplete. It can be shown that given any incomplete system, there are always square integrable functions f(x), g(x), . . . whose Fourier series do not converge to f(x), g(x), . . . in the mean. The reader will easily recognize that Bessel's inequality (7.1) is the analog of the inequality (10.7).

The case of a normabed system (10.6) corresponds to the case where the vectors i, j, L are unit vectors. In both cases, all the formulas simplify.

Page 77: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

PROBLEMS ORTHOC~ONAL S Y ~ M S 63

For example, the formulas (1 0.2) become

(where a, b, and c now coincide with the projections of the vector r on i, j, k, respectively), while the formulas for the Fourier coefficients become

The reader who is somewhat familiar with vectors in an n-dimensional space (in which case there are n pairwise orthogonal vectors) will find the analogy discussed hen even more natural. However, no matter how large n is, the transition from n orthogonal vectors to an infinite number of orthogonal vectors @ear in mind that we are treating an orthogonal system of functions as a system of vectors) cannot be regarded as simply a quanti- tative change; in fact, a qualitative change in behavior occurs, since instead of ordinary sums, we have to deal with infinite series and convergence in the mean.

PROBLEMS

1. Give another proof of the Schwan inequality (4.1) by considering the inequality

2, Prove the ineauality

where ai and bi are arbitrary real nwinbers. Comment. This result, known as the Cauchy tneqwltty, is the discrete

analog of the Schwarz inequality.

3. Let the polynomial P(x) = a0 + alx + + a,# have coefficients satisfying the relation

il

Prove that

Show that this inequality continues to hold if 4 2 is replaced by ,=/a.

4. Show that if nl, nz, . . ., nk are distinct nonzero integers, then

Page 78: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

64 ORTHOGONAL S Y ~ S CHAP. 2

Comment. It has been conjectured by the English mathematician J. E. ~ittlewood that there is a constant c such that the above integral 3 c log k. This remains unproved, although the integral has been shown to be 3 c (log k)lf4.

5. Let f and g be square integrable functions. Prove that

I f + I llfll + IrU* Comment. This result is often called the trkmgle inequality, since it is the

generalization of the familiar geometrical fact that the length of any side of a triangle is < the sum of the lengths of the other two sides.

6. Give an example of a sequence of functions which converges to 0 at each point of the interval [O,l], but which does not converge in the mean.

7. A system of functions cpo(x), ql(x),. . ., h(x), . . . which is not necessarily orthogonal is said to be complete if every square integrable function can be approximated in the mean by a linear combination of the cpl(x), i.e., if given any square integrable function g(x) and any s > 0, theie exist numbers ao, al, . . . , a,, such that

Show that if the system {cpl(x)) is complete, then any continuous function which is orthogonal to all the functions of the system must be zero. (Cf. the corm+ sponding result in Sec. 8 for the case of orthogonal systems.)

8. Let qo,ql,. . ., qn,. . . be a complete orthonormal system of functions. For which of the following systems is there no nonzero continuous function orthogonal to every function in the system:

a) 90 + 91990 + 92,cPo + 93, ; b) 'Po + 91, 91 + 92992 + 93, ; c) 90 + 2~r,c~1 + %2,92 + 2939.. .?

In Part c) we assume that the functions 9, are continuous and uniformly bounded, i.e., Ivn(x)l 6 Mfor a d x 6 b.

9. A system of functions cpdx), ql(x), . . ., cp,(x), . . . is said to be lhearIy independent if given any n, there is no set of numbers ao, at, . . . , a,, which are not all zero such that the linear combination aocpo(x) + alqr(x) + + anqn(x) is identically zero. Show that

a) An orthogonal system of functions is linearly independent; b) The functions 1, x, x2, x3,. . . are linearly independent.

10. Given a linearly independent system of functions f0, fi, . . . , &, . . . defined on the interval [a, b], we define a new system go, gl, . . . , gn, . . . as follows:

8 0 = fo, (fi, go)

gl fi - ,w 80,

Page 79: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

PROBLEMS ORTHOGONAL SYSTEMS 65

[Here (h g) denotes the scalar product of the two functions f and g, i.e.,

as in Sec. 10.1 This is the so-called Gram-Schmidt orthogonaIizatfon process. Interpret the process geometrically, and show that the new system go, gl, . . . , gn,. . . is orthogo~I and that llg. 112 # 0. Apply the process to the functions

thenby generating the Legendre polynomials (except for numerical factors). Show that a nonzero function is orthogonal to all the f i if and only if it is orthogonal to all the gt, and show that the system Cfi) is complete (set Rob. 7) if and only if the system (g,) is complete.

11. The Legendre polynomials are defined by the formula

Show that

J:, Pm(x)P,n(x) a5 = ifn = m.

12. Expand the following functions in Legendre polynomials:

0 for -1 < x < 0, O < x < 1;

Page 80: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

CONVERGENCE OF TRIGONOMETRIC FOURIER SERIES

I. A Consequence of Bessel's Inequality

For the basic trigonometric system 1, cos x, sin x, . . . , cos nx, sin nx, . . .

we have

Let f(x) be a square integrable function defined on the interval [-n, z]. Then, as applied to the system (1.1), Bessel's inequality (see Ch. 2, Sec. 6) becomes

Page 81: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 2

so that

From now on, we shall write Bessel's inequality for the basic trigonometric system in this form. It will be shown later (see Ch. 5, Sec. 3) that the equality actually holds in (1.2). However, for our present purposes, the inequality in (1.2) is sufficient.

The inequality (1.2) implies the following result concerning the conver- gence of the series in the right-hand side:

THEOREM. The sum of the squares of the Fourier coejfcients of any square integrable function always converges.

It is useful to note that this series always diverges for any other kind of function, i.e., for any function which is not square integrable. However, we shall not prove this fact.

2. The Limit as n + co of the Trigonometric Integrals

a x) cos nx dx and p f(x) sin nx dx f fc a

It is an immediate consequence of the preceding theorem that

lim a,, = lim b,, = 0 -00 -00

for any square integrable function, sina the general term of a convergent series must approach zero as n + a. But

a,, -- 1 f(x) cos nu d*, X -rr

b.. = Lr f(x) sin nx d*, X -n

and therefore

lim 1'1 f (x) cos nw dx = lim Jk f (x) sin nx dw = 0. (2.2) -00 -It *a0 -It

It follows from (2.2) that

lim Jb f (x) cos nx dw = lim Jb f (x) sin nx dx = 0, *a a -00 a

(2.3)

for any Interval [a, b] whatsoever. (We temporarily assume that f(x) is square integrable, although we shall drop this requirement soon.) To show

Page 82: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

68 CONVBROENCB OF T R I ~ ~ N O M B T R I ~ FOURIER SERIES CHAP. 3

this, suppose first that a < b ( a + 2n, i.e., b - a ( Zx, and set g(x) = f(x) for a ( x g b and g(x) = 0 for b < x ( a + 2x. The function g(x) is obviously square integrable on the interval [a, a + 2x1. We now extend g(x) periodically (with period 2x) over the whole x-axis. Then, by a pro- perty of periodic functions (see Ch. 1, Sec. 1)

In2= Ax) cos nx dx = rX g(x) cos nx dx, a

so that by (2.2) a+2n

lim I g(x) cos nx d* = lim g(x) cos nx dx = 0. n-+o0 a -=

On the other hand, it follows from the definition of the function g(x) that

r 2 = g ( x ) a a s nx d~ = [f(x) ws nx dr,

and therefore

The same argument applies to the second integral in (2.3). Finally, if b - a > 2x, then the interval [a, b] can be divided into a finite number of subintervals of length no greater than 2rr, for each of which the property summarized by (2.3) has already been proved. But this implies that the property also holds for the whole interval.

We now get rid of the requirement that f (x) be square integrable, and also of the requirement that n be an integer. To do this, we need two lemmas, which are rather obvious from a geometrical standpoint.

LEMMA 1. Let f(x) be continuous on the interval [a, b]. Then, for every E > 0, there exists a continuous, piecewise smooth fiMction g(x) such that

for all x (a ( x ( b).

Proo$ We divide the interval [a, b] into subintervals by the points

and for g(x) we take the continuous function for which g(xJ = f(x;) (k = 0,1,2, . . . , m) and which is linear on each interval [ x ~ - ~ , xk] (k = 1,2, . . . , m). The graph of the function y = g(x) is represented by a broken line with vertices on the curve y = f(x) (see Fig. 30). Ob- viously g(x) is a piecewise smooth function. Since f(x) is continuous,

Page 83: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

the subintervals into which [a, b] is subdivided can be chosen small enough to make (2.4) valid for all x in [a, b].

LEMMA 2. Let f (x) be absolutely integrable on the interval [a, b]. %n, for m y a > 0, there exikts a continuous, piecewice smooth function g(x) such that

ProoJl The function f (x) can have only a finite number of points of discontinuity, and in particular, only a finite number of points at which it becomes unbounded. We include every such point in an interval of such small length that the sum of the integrals of the function I f(x)l over these intervals does not exceed 4 3 . Next, define the auxi- liary function @(x) as being equal to f(x) outside these intervals and equal to zero inside them. @(x) is bounded and can have only a finite number of discontinuities, and obviously

Next, we include each point of discontinuity of O(x) in an interval of such small length that the total length 1 of all these new intervals satisfies the condition

where M is any number such that l@(x)l < M for a 4 x 4 b. Now, consider the continuous function F(x) which equals O(x)

1 This proof is based on the same idea as the proof of the formula (9.2) of Ch. 2.

Page 84: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

70 CONVEROBNCE OF TRIOONOMETRIC FOURIER SERIES CHAP. 3

outside the intervals just described and which is "linear" inside each of them (see Fig. 29 of Ch. 2). Obviously we have

Finally, according to Lemma 1, there exists a continuous, piecewise smooth function g(x) for which

I% I F(x) - dx) I < 3(b - a) (a < x < b),

so that

Then it follows from (26), (2.7), and (2.8) that

as was to be shown.

Remark. If f(x) is an absolutely integrable periodic function! then g(x) can be taken to be periodic.

THEo~rn.2 For any absolutely integrable function f(x), we have

lim [ /(x) cos mx dx = lim Jb f (x) sin mx dx = 0, (2.9) m+w m+ao a

where m is not necessariZy an integer.

Proof. Let E be an arbitrarily small positive number. By Lemma 2, there exists a continuous, piecewise smooth function g(x) such that

Consider the expression

I c f ( x ) cos mx dx I=IC mx) - g(x)] ms mx dx + Ib g(x) cos mx d l a

2 This result is usually known as the Riem-Lebesgue lemma, (~(utslator)

Page 85: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 2 CONVERGENCE OF TRIOONOMETRIC FOURIER SERIES 7 1

Integrating by parts, we obtain

b x=b g(x) cos tnx dx = 1. [g(x) sin mx] m x-a

- 1. Ib gt(x) sin mx dx. m o

The expression in brackets and the integral on the right are obviously bounded. Therefore, for sufficiently large m, we have

By (2.10) and (2.12), it follows from (2.1 1) that

Ir/(x) cos mx d*l< r a

for all sufficiently large m, i.e.,

lim Sb (x) cos mx dx = 0. m+oo a

The same argument applies to the second integral in (2.9), and the theorem is proved.

Recalling the formulas for the Fourier coefficients, we can phrase this theorem as follows: The Fourier coeflcients of any absolutely integrable function approach zero as n+ 00. At the beginning of this section, we proved this property for any square integrable function, and we have now extended it to the case of any absolutely integrable function. It should be noted that if we drop the requirement that the function be absolutely inte- grable, its Fourier coefficients may not converge to zero as n + 00.

3. Formula for the Sum of Cosines. Auxiliary Integrals

We now show that

) + cos u + cos 224 + + cos nu = sin (n + ))u 2 sin (~12)

4

To prove this, we denote the sum on the left by S. Obviously we have

U U 2s sin 5 = sin 3 + 2 cos u sin Z

Page 86: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

72 mNVEROENCE OF TRIC3ONOMETRlC FOURIER SERIES CHAP. 3

Applying the formula

2 cos a sin p = sin (a + p) - sin (a - p) to every product on the right, we obtain

U 2s sin - = sin + sin - u - sin i) + (sin u - sin - u + 2 Z " ( : 3 , 2

+ (sin (n + i ) u - sin (n - i)u) - sin (n + i)u. Therefore

S = sin (n + ))u 2 sin (up) '

as was to be proved. Next, we prove two more auxiliary formulas, Integrating the equality

(3.1) over the interval [-z, z ] and dividing the result by z, we obtain

1 $ sin (n + +)u I = - n -n 2 sin (2412)

for any n whateoever (since the integrals of the cosines vanish). It is easy to see that the integrand in (3.2) is even (since changing the sign of u changes the sign of both the numerator and denominator and leaves their ratio un- changed). Therefore we have

1 sin (n + +)u du 1 o 2 sin (~12) = 2'

4. 'The Integral Formula for the Partial Sum of a Fourier Series

Let f(x) have period 2z and suppose that 00

+ 2 (ak cos kx + bk sin kx). f(x) - & = I

Writing n

SAX) = + 2 (ak cos kx + bk sin kx), k= 1

and substituting the expressions for the Fourier coefficients, we obtain

+ $' f(t) sin kt dt sin kx -It I

Page 87: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 4 CONVEROENCE OF TRIGONOMETRIC FOURIER SERIES 73

n 1 n = i f(t) [I + (cos kt cos kx + sin kt sin k 3 dt

X -rc k== 1 I

or using formula (3.1)

3.O = l j:'f (0 sin [(n + +XI - x)l& x 2 sin [ f ( t - x) J

We now change variables by setting t - x = u. The result is lc-x

sn(x) = - sin (a + f . 1 ~ due x I -n-x f(x + 2 sin (~12)

Now, the functions f (x + u) and

sin (n + +)u 2 sin (2412)

are periodic in the variable u, with period k [see (3.111, and the interval [ -x - x, x - x] is of length k. Therefore, the integral over this interval is the same as the integral over the interval [-IT, ,n] (see Ch. 1, Sec. 1) and we obtain

1 % sin (n + +)u s~(x) = ; + 2 dn (@)

This integral formula for the partial sum of a Fourier series allows us to establish conditions under which the convergence of the series to f(x) can be guaranteed.

5. Rig ht-Hand and Left-Hand Derivatives

Suppoae that the function f(x) is continuous from the right at x, i.e., f(x + 0) = f(x). Then, we say that f (x) has a right-hand derivative at the point x if the limit

lim f ( x + u) - f 8 4 - 4 U

= f;(4

exists and is finite. If f(x) is continuous from the iefi at x, i.e., f (x - 0) = f(x), and if the limit

lim f ( x + U ) - f (4 A U

= fI(x)

exists and is finite, then we say that f(x) has a left-hand derivative at x.

Page 88: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

74 CONVBROENCE OF TRIOONOMETRIC FOURIER SERIES CHAP. 3

In the case where f:(x) = f:(x), the function f(x) obviously has an ordinary derivative at x, which is equal to the common value of the right- hand and left-hand derivatives at x, i.e., the curve y = f(x) has a tangent at the point with abcissa r In the case where f:(x) and fi(x) both exist but are unequal, the curve f(x) has a "comer" but we can still speak of right- hand and left-hand tangents (as indicated by the arrows in Fig. 31).

Now let x be a point where f(x) has a jump discontinuity. Then, if instead of (5.1), the limit

lim f(x + U) - f ( ~ + 0) = f:(X) IC*O U

exists and is finite, we again say that f(x) has a right-hand derivative at r Similarly, if instead of (5.2). the limit

exists and is finite, we say that f(x) has a left-hand derivative at r The existence of a right-hand derivative at a point of discontinuity x = xo is equivalent to the existence of a tangent at x = xo to the curve y - f+(x), equal to f(x) for x z xo and equal to f(xo + 0) for x = x,,, (Thus, the function f+(x) is defined only for x 3 xo.) In just the same way, the existence of a left-hand derivative at x = xo is equivalent to the existence of a tangent at x = xo to the curve y = f-(x), equal to f(x) for x < xo and equal to f(xo - 0) for x = xw (The function f-(x) is defined only for x ( xw)

As an example, consider the function

- 3 for x < 1 f(x) = for x = 1

t / i for x z I , shown in Fig. 32 This function has a jump discontinuity at x = 1, and obviously

f+(x) = .\Tx for x s 1, f-(x) = - x3 for x ( 1.

Page 89: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Therefore, we have

fi(1) = ( - 3 ~ 2 ) , ~ = -3.

The corresponding tangents are indicated by arrows in the figure.

6. A Sufficient Condition for Convergence of a Fourier Series at a Continuity Point

We now prove the following important

T~OREM. Let f (x) be an absoZute1y integrabk fMction of period 2%. Then, at every contfmrlty point where the right-hand and left-hand dkriva- fives exist, the Fourier series of f(x) converges to the value f(x). In particub, thi& is the case at m r y point where f (x) has a derivative.

Proof. Let x be a continuity point of f(x), where both the right- hand and left-hand derivatives exist We have to prove that

lim s&) = f ( ~ ) . -a0

By (4.1), this is equivalent to proving that

lim r f(x + u) sin (n + +)u 2 sin (2412) du = f(x).

,,+a0 % -= Moreover, since it follows from (33) that

sin (n + f)u h,

we can write sin (n + +)u du = 0, (6.2)

Page 90: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

76 CONVERGENCE OF TRIWNOMBTRIC FOURIER SERIE3 CHAP. 3

instead of (6.1). Thus the problem has been reduced to proving (6.2).

We begin by proving that the function

cp(u) = f ( x + u ) - f ( x ) - f ( x + u ) - - f ( x ) - U

2 sin (~12) u 2 sin ( ~ 1 2 ) 05-31

(x fixed) is absolutely integrable. Since f(x) has a right-hand and a left-hand derivative at the point x, the ratio

remains bounded as u+ 0. In other words, there exists a number 8 > 0 such that

I $ M = const

for - 8 ( u ( 6. For u # 0, this ratio has discontinuities only where f(x + u) is discontinuous. But f(x + u) is absolutely integrable [since f(x) is absolutely integrable], and hence can have only a finite number of discontinuities. Therefore, the function (6.4) is absolutely integrable on [- 6,6].

Outside [- 6, 61, the ratio (6.4) is absolutely integrable, being the product of the absolutely integrable function f(x + u) - f(x) and the bounded function l /u ( lul 3 6 implies that 1 1 1 ) Thus, finally, (6.4) is absolutely integrable both on [-8.61 and outside [- 6,6], and hence is absolutely integrable on [ -x, x].

On the other hand, the function U

2 sin (~12) is continuous for u # 0 and approaches 1 as u + O,3 and therefore is a bounded continuous function, which is undefined only for u = 0. Thus, the function cp(u) defined in (6.3) is absolutely integrable, being the product of the absolutely integrable function (6.4) and the bounded function (6.5).

Now that we have proved that p(u) is absolutely integrable, we note that

sin (n + +)u J:n mx + - ~ ( x ) J 2 sin (~121 du = 111 I t p(u) sin (n + +)u dic.

Then, the desired relation (6.2) follows by using (2.9).

3 Use the familiar formula lim sin or - a+O a = 1.

Page 91: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 7 CONVERGENCE OF TRIM-C FOURIER SERm 77

7. A Sufficient Condition for Convergence of a Fourier Series at a Point of Discontinuity

Next we prove the following

THEOREM. Let f (x) be an absolutely integrable f w i o n of period 2%. Then, at every point of discontinuity where f(x) has a right-hand and G left-hand derwative, the Fourier series of f(x) converges to the value

Proof. According to (4.1), we have to prove that

It is sufficient to show that I

We shall only give the proof of (7. l ) , since the argument leading to (7.2) is essentially the same.

According to (3.3)

f ( x + O ) 1 sin (n + +)u elk. 2 = n - rf o (x + O) 2 sin ( ~ 1 2 )

Therefore, we can prove the relation

instead of (7.1). First we prove the absolute integrability on the interval [0, n] of the following function of the variable u:

d u ) = f ( x + u ) - f ( x + O ) - f ( x + u ) - f ( x + O ) - u

2 sin (2412) U 2 sin ( 4 2 j Since f(x) has a right-hand derivative at x, the ratio

remains bounded as u + Of From this we conclude ljust as in the case

4 To prove (7.2), we have to consider f Cx + u) - fCT - 0) 6 < 0) U

instead of (7.4).

Page 92: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

78 CONVERGENCE OF TRIOONOME~RIC FOURIER SERIES CHAP. 3

of the ratio (6.4) of Sec. 6] that ( 7 4 is absolutely integrable on [0, x]. Then, since the function

u 2 sin (d2)

is bounded, the function cp(u) is absolutely integrable on [0, n]. But

A sin (n + +)u jo V(X + - + O)I sin (u12) du = I' 0 ~ ( u ) sin (n + +)u d ~ ,

and therefore the desired relation (7.3) follows by using (2.9).

8. Generalization of the Sufficient Conditions Proved in Secs. 6 and 7

An analysis of the proofs given in Secs. 6 and 7 leads to the conclusion that the existence of right-hand and left-hand derivatives at the point x was needed only to be able to prove the absolute integrability of the ratio

in Sec. 6 [see (6.4) et seq.] and the absolute integrability of the ratios

in Sec. 7 [see (7.4) et seq.], where x is fixed and the ratios are regarded as functions of u. Therefore, if we ussume this absolute integrability (relin- quishing the requirement that the right-hand and left-hand derivatives exist), we obtain the following more general convergence criterion:

T H ~ R E M . The Fowfer series of an absolutely integrable function f(x) of period Zx converges to f(x) at every point of continuity for which the ratio (8.1) is an absolutely integrable fMcrion of the variable u, and converges to the value

at every point of discontinuity for which both ratios (8.2) are absolutely integrable.

Page 93: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 9 CONVERGENCE OF TRIGONOMETRIC FOURIER SERIES 79

9. Convergence of the Fourier Series of a Piecewise Smooth Function (Continuous or Discontinuous)

The following theorem is a consequence of the results of Secs. 6 and 7:

'I'HE?OREM. If f(x) is an abso~utely integrable funcion of period 27c which &piecewise smooth on the interval [a, b], then for all x in a < x < b, the Fourier series off (x) converges to f(x) at points of continuity and to the value

at pobts of discontinuity. (The convergence may fail at x = a and x = b.)

Proof. The theorem is a simple consequence of the fact that a piecewise smooth function bn [a, b] (see Ch. 1, Sec. 9) must have a right-hand and a left-hand derivative for every x in a < x < b, so that we need only apply the theorems of Secs. 6 and 7. This is obvious at the points where f(x) has a derivative. At the corners of f(x), we use L'Hospital's rule and obtain

lim f (x + - f ( 4 U--*O U

= lim f (0 = f (x + O), U--*O

since x < c < x + u, so that t + x, k z x. Similarly, at a point of discontinuity off (x), we have

lim f(x + -f(x + 0) U

= lim f'(6) = f (x + 0). u+o u+O

In other words, f (x) has a right-hand derivative both at a corner and at a point of discontinuity. The existence of the left-hand derivative is proved in the same way.

As for the end points of the interval [a, b], the conditions of the theorem imply only that the right-hand derivative exists at x = a and that the left-hand derivative exists at x = b. Therefore, the criteria of Secs. 6 and 7 cannot be applied at these points. However, if the inter- val [a, b] is of length 2x, then it is easily seen that f(x) is piecewise smooth on the whole x-axis, since f(x) is periodic. In this case, the Fourier series converges everywhere.

Thus, finally, we have proved the first part of the criterion formulated in Ch. 1, Sec. 10. The second part of the criterion, pertaining to the absolute and uniform convergence of the Fourier series in the case where f(x) is continuous, will be proved in the next section.

Page 94: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

80 ~NVEROBNCE OF TRIGONOMETRIC FOURIER SERIES CHAP. 3

10. Absolute and Uniform Convergence of the Fourier Series of a Continuous, Piecewise Smooth Function of Period 2%

Let f (x) be a continuous, piecewise smooth function of period 2%. Then, the derivative f (x) exists everywhere except at the corners of f (x) and is a bounded function (see Ch. 1, Sec. 9). Therefore, applying the formula for integration by parts (which is permissible because of Ch. 1, Sec. 4), we obtain

1 a,, = - E f ( x ) cos nx d*

X

1 X a R

= - I f (x) sin n*] 7Cn

1 = = ; sin nx

1 %=%

= - - V ( x ) cos nx] 1

ma m + ~ l J - r j"(x) cos nx d . ~ .

The terms in brackets in the right-hand side of both formulas vanish. Thus, denoting the Fourier coefficients of the function f '(x) by d, and b;, we find that

Since f (x) is bounded and hence square integrable, it follows from the theorem of Sec. 1 that the series

converges. Next, consider the obvious inequalfies

which imply

On the right appears the genera1 term of a convergent series. Therefore the series

Page 95: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

also converges. But then it follows from (10.1) that for any continuow, piecewise smooth function, the series

converges. Remark. To prove the convergence of the series (10.3), we used only

the convergence of the series (10.2); this series always converges whenf'(x) is square integrable v(x) may not exist at certain pointss]. Therefore, (10.3) also converges in this case [when f(x) is a continuous function of period 2n]. We now consider a very simple but very important fact. Suppose we are

given a trigonometric series 00

9 + 2 (an cos nx + bn sin nx) 2 n=l

which is not assumed in advance to be the Fourier series of any function. Then the following result holds :

THEOREM 1. If the series

converges, then the series (10.4) converges absolutely and unforrnk'y, and therefore has a continuous. sum of which it is the Fourier series (see Theorem 1 of Ch. 1, Sec. 6).

Proof. Since

lan cos nx + b, sin nxl 6 )a,, cos nxl + 1 bn sin nxl

the terms of the trigonometric series (10.4) do not exceed the terms of the convergent numerical series (10.5). Then, the theorem follows from Weierstrass' M-test (Ch. 1, Sec. 4).

This theorem and the theorem of Sec. 9 imply the following

THBOREM 2. The Fourier series of a continuous, piecewise smooth function f (x) of period 2% converges to f (x) absolutely and uniformly.

This proves the second part of the convergence criterion of Ch. 1, Sec. 10. It follows from this theorem that a continuous, piecewise smooth ftnction

5 I.e., f'(x) may not exist at a finite number of points in each period.

Page 96: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

82 CONVBROENCB OF moo~onmnuc muwm SBRIBS CHAP. 3

f(x) of period Zx can be well approximated for large n by the partial sums s,,(x) of its Fourier series; in fact, this is just what uniform convergence means! (See Ch. 1, Sec. 4,)

As an illustration, consider the periodic, continuous, piecewise smooth function f (x) equal to 1x1 for -x < x ( 7c. In Example 2 of Ch. 1, Sec. 13, we proved that

cos 3x cos 5x 4 ( m a x + - f(x) = - - - + 52

+ . . .). 2 7c 32

Figure 33 shows the graph of f(x) and of the partial sum

s5(x) = 2 - ;; (cos X + jt cos 3x + wF)o

of its Fourier series, We see that the two curves are already quite close to each other for n = 5.

The remark preceding Theorem 1 allows us to state the following general- ization of Theorem 2:

THEOREM 3. The Fourier series of a continuous fmction f(x) of period 27c, whose derivative (which may not exist at certain points) Is square integrable, converges absolutely and unCormly to f(x).

I I. Uniform Convergence of the Fourier Series of a Continuous Function of Period 2n with an Absolutely Integrable Derivative

LEMMA 1. Let f(x) be a continuous function of period 2x, which has an absolutely integrable derivative (that may not exist at certain points),

Page 97: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

and let o(u) (a < u ( p) be a fMctfon with a continuous derivative. Tien, for any t > 0 whatsoever, the hequality

(j' /(x + U)O(Y) sin mu d~ < a I (11.1)

holh for all values of x, provided that m (which fs not necessarily an Integer) is mflciently large.

Proof. Integration by parts gives

1 -0 jhP f(x + u)&) sin mu du = - [- f(x + v)o(u) cos mu] m UPQ

(1 1.2)

The term in brackets is obviously bounded. Since

V(x + u)o(u)r = f (x + u)o(u) + f (x + u)08(u), (1 1.3)

it is easily seen that the integral on the right is also bounded. In fact, o(u) and flx + u)ot(u) an bounded, i.e., have absolute values which do not exceed some constant M. But then, by (1 1.3)

where we have used the fact that f(x), and hence lf'(x)l, is periodic. We have also assumed that p - a < b, which is not essential but is sufficient for our purposes.

Now that we have shown that the term in brackets and the integral in (1 1.2) are bounded, the validity of (1 1.1) is obvious.

LEMMA 2. The integral sin mt ' ' 10 2 sin (t12) "

fs bounded for -IT < u d x and any m.

Proof. We have

I = ji sin mt dt + r o(t) sin mt dt, t 0

where 1 1

= 2 sin (t/2) - i*

Page 98: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

84 CONVERGENCE OF TRIQONOMBTIUC FOURIER SERLES CHAP. 3

Applying L'Hospital's rule, we see that a(?) and a'(?) are continuous [if we set o(0) = 0 and o'(0) = A]. The ~econd integral in (11.5) is obviously bounded. On the other hand, setting mt = x, we obtain

mu sin x = -&, 0 X

and it is easily seen that the last integral does not exceed the area of the first "hump" of the curve y = sin x/x (see Fig. 34). Therefore, both integrals on the right in (11.5) are bounded, i.e., the integral I is bounded.

THW)REM= me Fourier series of a continuous fiction f(x) of perwd 2n with an absolutely integrable derivative (which may not exist at certain points) convegar ~IforrnJy to f(x) for all r

Proof. Consider the difference

sin mu 4 (1l.a)

already calculated in Sec. 6, where we have set m = n + 4. Choose any e > 0 whatsoever, and let 8 be a number between 0 and n. Then divide the integral in (1 1.6) into three integrals 11, 12, I, over the inter- vals [- $,a], [a, x], [-=, - 83, respectively. Integration by parts gives

sin mt o 2 sin (t/2)

sin mt o 2 sin (t/2) dt] dw

For the value of the first term on the right we obtain [using the evenness of the function (sin mt)/2 sin (t/2)]

8 sin mt [ ~ ( x + 8 - /(x)) + (I(X - 8) - ~ ( ~ 1 1 I )lo 2 sin (t/Z) ''9

Page 99: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

which shows that this term does not exceed el2 in absolute value for all sofliciently small 6 [since f (x) is continuous and the integral is bounded, by Lemma 21. On the other hand, by Lemma 2

sin mt o 2 sin (t/2)

where M is a constant, for all sufficiently small 8, since the integral

is the increment of the continuous function

(x0 fixed) and is therefore small when 6 is small.6 Thus, finally

1111 a e,

for any x whatsoever, provided that the number 8 is chosen to be sufficiently small.

Furthermore

sin mu sin mu 1121 IJ':f(x + u) 2 (u12) + 1 T f ( x ) i-I dul r

for all x, provided that n is sufficiently large. This follows from Lemma 1, if we set

1 = 2 sin ( ~12 )

and a = 8, = x. A similar inequality is obtained for the integral 4. Combining results, we have

for all x, provided that n is sufficiently large.

12. Generalization of the Results of Sec. I I

What can we say about the nature of the convergence of a Fourier series if f(x) is continuous and has an absolutely integrable derivative not

6 Without loss of generality, we can ass- that -x g x 6 x.

Page 100: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

86 CONVEROENCE OF TRIOONOMETR~C FOURIER SERIES CHAP. 3

everywhere but only in some interval? We shall now concern ourselves with this question. First we strengthen Lemma 1 of Sec. 11.

LEMMA. Let f(x) be an absolutely integrable function of period 2rr, and let o(u) (a < u 6 p) be a fmtion with a continuous derivative. Then, for any r > 0 whatsoever, the inequality

( Jap f (x + u)o(u) sin mu du < r, I ( 1 2.1)

holdr for all x, ptovided that m (which is not necessarily an integer) h mflciently lmge.

Proof. Let lo(u)l M, M = const, and choose a continuous, piecewise smooth function g(x) of period 2% which satisfies the inequality

s I:, If(x) - g(x)l a < (see Lemma 2 of Sec. 2 and the remark). Then we have

I l:f(x + u)o(u) sin mu du I = ( j" ~ f ( x + u) - g(x + u)b(u) sin rnu L

+ I p g(x + u)o(u) sin mu dU < Ilf(x + I() - g(x + u)Wu)I (ik a I I + I J: g(x + u)(u) sin mu du . I (1 2.2)

If m is large enough, then by Lemma 1 of Sec. 1 I, the last integral does not exceed r/2. On the other hand

where we have used the periodicity of the difference f(x) - g(x) and have assumed that p - a 2 7 ~ Thus, (12.2) implies (12.1).

THEOREM. Let f (x) be an absolutely integrable fmction of period 2n, which ts continuous and has an absolutely integrable derivative on some interval [a, b]. (The derivative may not exist at certain points.) men, the Fourier series of f(x) converges e m l y to f(x) on every interval [a + 8, b - 81 (8 > 0).

Proof. I f the length of the interval [a, b] is not less than 275 it is clear that f(x) is continuous for all x and has an absolutely integrable derivative, so that by the theorem of Sec. 1 1, the convergence of its

Page 101: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 12 CONVERGENCE OF TRIOONOMETRIC FOURIER SERIES 87

Fourier series is uniform on the whole x-axis. Thus, we assume that the length of [a, b ] is less than 2n.

We introduce an auxiliary continuous function F(x) of period 2x, which equals f(x) for a ( x ( b, equals f(a) for x = a + 2n, and is "linear" on the interval [b, a + %] (see Fig. 35). Outside [a, a + 2n], the values of F(x) are obtained by periodic extension. It is easy to see that F(x) has an absolutely integrable derivative.

Next, we set @(x) = f(x) - F(x). This function is absolutely integrable and @(x) = 0 for a < x ( b. Obviously

= F(x) + and

sin mu

1 sin mu + I-= [qx + +) - qx)J 2 sin ( ~ 1 2 ) du (12.3)

where we have set m = n + +: Let c > 0 be arbitrary. By the theorem of Sac. 11, the Fourier series of F(x) converges uniformly to F(x), so that

for all x, provided that n is sufficiently large. Now let a + 6 < x ( b - 6. Then @(x) = 0 and therefore

l r c sin mu b = , LIt 'D(x + 2 sin (u/2) du.

~f -6 ( u ( 6, we have

a < x + u G b

Page 102: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

CHAP. 3

for the values of x under consideration, and hence

@(x + u) = 0.

Therefore

1 - S sin mu sin mu '2 = ; I-% ' ( x + U) 2 sin ( ~12 ) + 7t ' ~ W X 8 + u) sin (u,2) dl4

and it remains only to apply the lemma just proved to each of these integrals. As a result, we obtain

for a + 6 ( x ( b - 6 and all sufficiently lare n. Finauy, according to (12.3). it follows from (12.4) and (12.5) that

Isn(x) - f (x) I 1111 + 1121 ( 0,

for all x in the interval [a + 6, b - 61, provided that n is sufficiently large. This proves the theorem.

Rcmmk. In particular, the theorem is valid for an absolutely in* grable function f(x) of period 2% which is continuous and piecewise smooth on the interval [a, b].

To illustrate this theorem consider the piecewise smooth, even, periodic function f(x) which equals n 4 for 0 < x < x and -n/4 for -n < x < 0. In Example 5 of Ch. 1, Sec. 13, it was shown that

sin 3x sin 5x sin 7x f(x) = sin x + -

+ 5 + 7 +. . .

3

for x # kx, while f ( h ) = 0 (k = 0, f 1, f 2,. . .). In Fig. 36 we show the graph off(x) and of the following partial sums of

its Fourier series: s1(x) = sin X,

sin 3x s3(x) = sin x + 3

sin 3x sin 5x s5(x) = sin x + - 3 + 5 9

sin 3x sin 5x sin 7x SAX) = sin x + - + 5 + 7 ' 3

The figure clearly shows the uniform character of the approximation of the partial sums to f(x) on the intervals [ -x + 6, - 61 and [6, x - 81 (6 > O), on which f(x) is a smooth function. Note that the number 6 can be chosen

Page 103: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

to be arbitrarily small (but not equal to zero). It is easily seen that as 6 is made smaller, we need to take partial sums with higher indices in order to have a good approximation to f(x) on the intervals just mentioned (more precisely, to approximate f (x) with a specified degree of accuracy).

Page 104: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

90 -0ENCE OF T R I Q O N O ~ C F O ~ SERIES CHAP. 3

13. The Localization Principle

A change in the values of a function even on a small interval can change its Fourier coefficients considerably. However, if an absolutely integrable function/(x) has a right-hand and a left-hand derivative at the point x, or is continuous and has an absolutely integrable derivative in a neighborhood of x, then, according to Secs. 6,7, and 12, its Fourier series remains convergent no matter how the values of/@) are changed outside a neighborhood of x. This fact is a special case of the following proposition, known as the localization principle :

THEOREM. me behavior of the Fowler series of m, absolutely We- gr& function f(x) at the point x &pen& only on the o a k s of x in an arbitrarily small neighborhood of r

This proves that if the Fourier series of f(x) is originally convergent at x, then no matter how we change the values of the function (while leaving it absolutely integrable) outside a neighborhood (however small) of x, the series remains convergent, while if the series was originally divergent at x, it remains divergent.

Proof: We use the integral formula for the partial sums (sa Sec 4):

sin mu ahr

sin mu dk + ll + 12,

where we have set m = n + 3. Here 8 is an arbitrarily small positive number, and 11, 12 are the integrals over the intervals [a, x] and [-z, - 81 respectively. On these intervals, the function

1 2 sin (2412)

is continuous (since lul 3 8), and therefore, the function

is absolutely integrable. But then according to (2.9), the integral

approaches zero as m + 00. The same is true of 12. Thus, whether or

Page 105: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

not the partial sums of the Fourier series have a limit at the point x depends on the behavior as m + a, of the integral

1 8 sin mu ii 1- , f (~ + 2 sin (u/2) dU'

which involves only the values of the function f(x) in the neighborhood [x - 8, x + 81 of the point r This proves the localization principle.

14. Examples of Fourier Series Expansions of Unbounded Functions

Example 1. Let f(x) = -In 12 sin (xl2)J .7 This function is even and becomes infinite at x = 2kn (k = 0, f I , f 2, . . .). Moreover, f(x) is periodic, since

so that

ln 12 sin * : 2"1 = In 12 sin f 1. The graph of f(x) is shown in Fig. 37.

To prove that f(x) is integrable, it is sufficient to prove that it is integrable on the interval [0, z /3 ] (see the graph off (x)). Clearly, we have

- j"13 in 12 sin dx = - Isn" ln (2 sin 4) dx

= - [x ln ( 2 sin f)] xn/3 + J.nl3 x cos ( ~ 1 2 ) dw X-t e 2 sin ( ~ 1 2 )

= s In 2 sin - =I3 x cos (x/2) ( ) + 2 sin dx,

7 In x denotes the nutural logarithm of x. (Tr(u1slator)

Page 106: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

92 CONVBROBNCB OF TRIOONOMBTRIC FOURIER SERIES CHAP. 3

where we have dropped the absolute value sign, since 2 sin (~12) > 1 for 0 < x < 743. As r + 0, the quantity r ln (2 sin (~12)) approaches zero, as can easily be verified by using L'Hospital's rule, while the last integral converges to the integral

I 4 3 x cos (xp) o 2 sin (42) h,

which obviously has meaning, since the integrand is bounded.* Thus

I L ~ r I 3 in sin 1 dx &+O 8

exists, i.e., f(x) is integrable on the interval [O, x/3]. Moreover, Ax) is absolutely integrable on the interval [O, x/3], since it does not change sign there (soe Fig. 37).

Since f(x) is even, we have

First of all, we calculate the integral

We denote the last integral by Y and make the substitution x = 2t:

The substitution t = x - u gives

Therefore Y = x ln 2 + 2Y, so that

Y = --=In 2.

This implies that I = 0, so that a. = 0.

8 Recall that X 3 2 dn (42) = 1.

Page 107: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 14 CONVBRGENCB OF TRIGONOMElWC FOURIER SBRIBS 93

Furthermore, integration by parts gives an = - 2 ([In (2 sin (42)) sin nx Fm sin nx cos (~12)

lr n ] mo - k l o 2 sin (~12) 4 1 sin nx cos (~12) = -

m o sin (XI~)

(The first term in braces vanishes, since the indeterminacy for x+ 0 is easily "removed" by using L'Hospital's rule.) But

x 1 sin nx cos - = [sin (n + +)x + sin (n - +)XI, 2 Z and therefore

' I sin (n + 4)x dx + 1 in sin (n - flx a, 3 - m o 2 sin (~12) m o 2 sin (42) dx,

which by formula (3.3) of Sec. 3 gives

Since the function f(x) is obviously differentiable for x # 2h (k = 0, f 1, * 2,. . .), it follows from the theorem of Sec. 6 that

for x # Z ~ I C (k = 0, f 1, 2, . . .). It should be noted that for x = Zkrc, both sides of (14.1) become infinite. Thus, in this sense, the equation (14.1) can be regarded as valid for all x.

Setting x = x in (14.1), we obtain the familiar formula

Example 2. Let f (x) = In (2 cos (x/2)(. This function is even and goes b --a for x=(Zk+l ) r r [k=O, f 1 , f %...I. Setting x = t - n , we see that

i.e., the graph of the function ln 12 cos (~12)) is obtained from that of the function In 12 sin (x/2)l by shifting it by an amount r To obtain the Fourier series of f(x), it is sufficient to make fhe substitution t = x + x in the expansion

[see (14. I)]. The result is

I % cos 2x cos 3x In 2cos- s c o s x - - + - - . . a

2 3

Page 108: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

94 -0ENCE OF TRIOONOMETIUC FOURIER SERIES CHAP. 3

for x # (2 + 1)z [k = 0, f I, f 2, . . .I. Moreover, since both sides of (14.2) go to - ao for x = (2k + l)z, this equation can be regarded as valid for all x.

15. A Remark Concerning Functions of Period 21

In discussing questions of theory in this and later chapters, we do not talk about Fourier expansions of functions of period T = 21. However, the reader who is familiar with the contents of Chs. 1 and 2 will have no trouble in making the transition from the case of the "standard" period 2x to the case of any period.

PROBLEMS

1. Derive formula (3.1) from the formula

2. For each of the following functions, find the right-hand derivative at zerc. f;(O), if it exists, and find lim f'(x) = f'(0 +), if this limit exists:

x+o x>O

1 b) f (x) = X* sin - x '

1 c) Ax) = x3 sin -*

X

[In each case, f(0) = 0.1

3. Show that the theorem of Sec. 6 can be generalized in the following way: If f(x) is an absolutely integrable function of period Zx, if f(x) is continuous at the point xo, and if there are numbers c > 0, a > 0 such that

for all x in some neighborhood of xo (cf, Ch. 1, Prob. 7), then the Fourier series off (x) converges to the value f (xo) at the point x*

4. Let f(x), and g(x) be absolutely integrable functions with period 2x, whose Fourier series are

Page 109: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

PROBLEMS

and let

CONVERGENCE OF TRIOONOMETRIC FOURIER SERIES 95

h(x) = I k f ( x - t)g(t) dt, 2rs 0

Show that

and that Cn = a,&,,. In particular, if f(x) and g(x) are square integrable, show that

in., that the Fouriex series of Mx) is absolutely convergant.

5. Show that the Fourier series

can be written in the form

wh- pn = l/a: + b:. Express On in tams of an and bn.

6. Suppose that 9, . 0 and that

for a < x < b. Show that

Hint. Integrate equation (I) from a to 6, use the fact that

I c a (a + OJl 3 C O S ~ (m + en) = f + f cos k w 2% - f sin tUI sin 28, and then use equation (2.9).

7. According to equation (1 3.7) of Ch. 1, the function

has the Fourier expansion

Page 110: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

CHAP. 3

Let sn(x) be the nth partial sum of the series, i.e.,

and let

sin (n + +)x Dn(X) ' 2 sin (r/~)

Show that

x sin nt nX sin t b) I Dn(*) clt = I - dt + wn(x) I. dt + u.(x),

0 0 t

8. Using the notation of the preceding problem, show that

sin t A ~ ~ s ~ ( ~ ) = ~ ~ ~ > j ~ + ~ = ~ . n+a

Comment. Thus, near x = 0 (a point of discontinuity), the partial sums of the Fourier Series of the function f(x) of the preceding problem exceed the value of the function at x = 0 by the amount

This illustrates the so-called Gibbs phenomenon, according to which the Fourier series of a discontinuous function "overshoots" its limiting value at a point of jump discontinuity.

Page 111: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

TRIGONOMETRIC SERIES WITH

DECREASING COEFFICIENTS

. Abel's Lemma

The following result will be needed below: ABEL'S LEMMA. Let

be a numerical series (with real or complex terms), whose partial sums a, satisfy the condition

141 M, where A4 is a constant. Then, if the positive numbers ao, a,, a?, . . . , an, . . . approach zero monotonically, the series

converges, and its sum s satisfies the inequality

Is1 < M a o Proof. Let

then, since uo = a0 and un = a, - a,-1 (n = 1,2, . . .), we have

Page 112: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

CHAP. 4

It follows that

Now consider the series

- at) + al(al - ai) + . + q l ( b l - a,,) + (1 *4)

This series converges, since the absolute values of its terms do not exceed the comsponding terms of the following convergent series with nome&ative terms :

M(% - al) + M(al - ad + + M(akl - a,,) + = M ( ~ - a l + a l - a 2 + - o - + ( 5 r l - a n + * * * ) = Ma&

The right-hand side of (1.3) is the nth partial sum of the series (1.4), and therefore, as n + oo, it approaches a definite limit, whose absolute value does not exceed the number Ma* But then the left-hand side of (1.3) also approaches a limit as n + a, and

Since

we have

lim a,,a,, = 0. I)-*oO

Therefore

lim Sn = s *a0

exists, i.e., the series (1.1) converges, and s satisfies the inequality (1.2).

2. Formula for the Sum of Sines. Auxiliary Inequalities

We now prove the formula

sin x + sin 2x + . + sin nx = cos ( ~ 1 2 ) -- cos (n + +)x. (2.1) 2 sin (~12)

Page 113: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

(*s?sswu! u ss papunoq IOU s! auay prnr '? + u X~sno!~qo s! tuns ay) 'w = x rod) I)(- '2 '1 7 '0 = y) # x JOJ

('ppunoq osp s! muay pus saqsruon mns a q 'uyl: = x rod) *x paxg Xra~a JOJ papunoq s! saqs JO mns aq, s ~ o q s s m *un # x roj

Page 114: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

100 TRIGONOMETRIC SERIES WITH DECREASINO COWFICIEN'IS CHAP. 4

3. Convergence of Trigonometric Series with Monoton ically Decreasing Coefficients

Consider the two trigonometric series

2 bn sin nr,

which we do not even assume to be the Fourier series of any functions.

THEOREM 1. If the coeficients an and bn are positive and decrease monotonically to zero as n + oo,1 then the series (3.1) and (3.2) converge for any X, except possibly the a u k s x = 2kz (k = 0, f 1, & 2, . . .) in the case of the series (3.1).

Proof; If the sums of the coefficients a,, and b,, converge, then the theorem follows from Theorem 1 of Ch. 3, Sec. 10. In the general case, consider the series

f + cosx + cask +-+ cosnx + - a , (3.3)

whose partial sums an(x) are bounded for any x # 2lnr. Then, to prove the theorem for the series (3.1), we apply Abel's lemma. By (2.2)s the same argument applies to the series (3.2).

Remark. Of course, Theorem 1 and the theorems which follow remain valid if the requirement that the coefficients be nonincreasing is satisfied not for all n, but only starting from a certain value of n. In particular, the theorem is true in the case where some of the early coefficients vanish. For example, the series

4% COS IZX

converges for x # Zkn. For x = 2kn, this series becomes

and hence diverges.

We now make Theorem 1 more precise.

1 La, a1 3 a2 bt 3 and lim an = limbn = 0. (Zhslator) n--* OD n+oo

Page 115: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 3 TRIGONOMETRIC SERIES WITH DECREASING COEFFICIEN'I3 10 1

THEOREM 2. the coeflcients an and bn are positive and decrease monotonicalZy to zero as n -, 00, then the series (3.1) and (3.2) converge uniformly on any interval [a, b] which does not contain points of the form x = & ( k = O , k1, k 2 ,... ).

Proof: If the sums of the coefficients an and bn converge, then (3.1) and (3.2) converge uniformly on the whole x-axis by Theorem 1 of Ch. 3, Sec. 10. In the general case, since the sums of the series (3.1) and (3.2) are periodic functions, it is suflicient to prove the theorem for every interval [a, b] contained in the interval [O, 2z]. The proof is the same for both series, and therefore we confine ourselves to the case of the series (3.1).

Let o > 0 be arbitrary. For u ( x ( b, we consider the remainder

S(X) - s~(x) = a,,+1 cos (n + 1)x + a,* cos (n + 2)x + * * * (3.4)

of the series (3.1), and apply Abel's lemma. To do this, we set

where OJX) and a,,+,,,() are partial sums of the series (3.3). Then by (2.31,

Since 0 < a < x 6 b < 27t, we have

X sin - 2 p > 0, 2

where p is the smaller of the numbers sin (a/2) and sin (b/2) (see Fig. 38).

M o r e , setting M = 1/p, we find

Irm(x)l < M cas t

for all x in the interval [a, b]. Thus, Abel's lemma is applicable to the

Page 116: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

numbers a*,, a,%. . ., . . and to the series (3.4), ar gives

IJ(x) - sn(x)I g man-^-1 for any x in the interval [a, b]. Since an + 0 as n + m. have

Ma,,+l < E

for all sufficiently large n. In other words, for ouff ciently large n and for any x in [a, b], the inequality

holds, which proves the uniform convergence of the series (3.1).

Theorems 1 and 2 imply

THEOREM 3. If the coeflcier?~ an and b, are positive and decrease monotonically to zero as n + a, then the j i i rtions

f (x) = + 2 an cos nx, go = 2 bn sin nx n u 1 n o 1

of period 2z are continuous for all x, except possibly for the values x =Zkx (k=O, &1, &2 ,... ).

Proof. If the sums of the coefficients an and bn converge, then the theorem follows from Theorem 1 of Ch. 3, Sec. 10. In the general case, we can include any point xo # Unr in an interval [a, b] which does not contain points d the form x= 2kz. In this interval, . the two series converge uniformly (by Theorem 2) and hence their sums are continuous (see Ch. I, Sec. 4). In particular, they are continuous at x = xo. Since 3.-J is any point different from a point of the form x = 2h, the tileorem is proved.

As an illustration, consider the expressions

Page 117: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

vith which we are already familiar [see Example 1 of Ch. 3, Ssc. 14 and formula (13.7) of Ch. 11. The graph of f(x) is shown in Fig. 37, and the graph of g(x), together with its periodic extension, is shown in Fig. 39.

*4. Some Consequences of the Theorems of Sec. 3

We now present some interesting consequences of the theorems proved in Sec. 3.

THEOREM 1. If the coeflcients an and bn are positive and decrease monotonically to zero ar n + m, then the serks

00

2 ( - lrb, sin nx n= 1

have the followihg properties:

1 ) They converge for all values of x except possibly at the points x = (2 + 1)n [k = 0, f 1, f2,. . .] in the case of the series (4* 1 ) ;

2) m e convergence is uniform on every mterval [a, b] not contabhg these points;

3) The sums of the series are continuous, except possibly at these points.

Proof. We substitute x = t - x in (3.1) and (3.2), thereby obtaining the series

QO 00

9 + 2 an cos n(t - X ) = + 2 ~,,[cos m cos nt + sin m sin nt] 2 n=l n= I

00 00 2 bn sin n(t - x) = 2 bn[cos m sin nt - sin m cos nt] n=l n=l

These are alternating series to which Theorems 1, 2, and 3 of Sec. 3 apply, provided the points t = x + 2 h = (2k + l)x now play the role of the points x = 2 h .

Page 118: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

104 TRIGONOMBTIUC SERIES WITH DECREASINO COEFFICIB~ CHAP. 4

Theorem 1 obviously remains valid if in (4.1) and (4.2). we write (- lF+* instead of (- 1)". To illustrate Theorem 1, consider the series

I ;I cos2x + cos3x 1n 2 cos - = COS X - -... 2 3 and

x sin 2x sin 3x Z = sinx - - + - -...

2 3 ( - x < X < n)

with which we are alnady familiar [see Example 2 of Ch. 3, Sec. 14 and formula (13.9) of Ch. 11.

The theorem just proved can be generalized even further. In fact, consider the series of the form

+ a,,+l cos@ + nm)x +.-, b, sinpx + b2 sin(p + m)x + b3 sin (p + 2m)x + (4.3)

+ bM1 sin (p + nm)x + 0 ,

where p and m are any numbers, and the coefficients an and bn are positive and converge monotonica11y to zero. In both of these series, the coefficients of x form an arithmetic progression with difference m. The following are examples of series of this type:

cos 5x cos 9x cos 13x coax+

+ 3 + 4 +... (p = l,m = 4),

(4.4) \ ,

sin 2x sin 5x sin 8x sin l l x - +- +- + Ins +...

ln 2 In 3 In 4 (p = 2,m = 3).

Observing that

cos (p + nm)x = cos px cos nrnx - sin px sin nmx, sin (p + nm)x = sin px cos nrnx + cos px sin nmx,

we can rewrite the series (4.3) as ~-

COS px 2 a,l cos nrnx - sin px 2 a , ~ sin nmx, n u 0 n=O

sinpx 2 bMl cos ~ n x + cospx 2 b,,+l sin nmx. n = O n=O

If we now set mx = t or x = tlm, we obtain a0 QD

cos 2 a** cos nt - s i n e 2 aMl sin nt, n = o n o 0

00 QD

s i n e 2 bn+l ~ s n t - C O S ~ 2 bMl sinnt. n u 0 n=O

Page 119: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Theorems 1, 2, and 3 of Sec. 3 are applicable to all four series appearing in (4.5). It follows that for anyp, the series (4.5) converge and have continuous sums for all t, except possibly values of the form t = Zkx, and that the convergence of the series (4.5) is uniform on any interval which does not contain these values. Thus, returning to the variable x, we have

THEOREM 2. If the coeflcients a, and b, are positive and decrease monotonicall) to zero as n + a, then the series (4.3) converge and have continuous surns for all values of x, except pmibly the oaks x = 2kxIm (k = 0, & 1, 5 . . .), and the convergence of the series is muorm on any interval not containing these points.

For example, for the first of the series (4.4), the "exceptional" points are x = %I4 = h l 2 and for the second series, they are x = 2h/3.

In the same way as we obtained Theorem 2, we can prove

THEOREM 3. the coeficients a,, and bn are positive and &crease monotonically to zero as n + a, then the series of the form

al cospx - a2cos(p + m)x + a 3 m @ + 2m)x -=-• , bl sin px - b2 sin ( p + m)x + b3 sin (p + 2m)x -

converge and have contirtwus sums for all x, except possibly

a d the convergence is unvorm on any interval not containing these points.

In the applications, the following theorem, which we cite without proof, is often useful:

THEOREM 4. With the hypotheses.of the preceding theorems, Y the series

converge, then the corresponding trigonometric series deBne absohtely integrablefiutctions (and are therefore the Fourier series of these functions). (See Theorem 2 of Ch. 1, Sec. 6.)

5. Applications of Functions of a Complex Variable to the Evaluation of Certain Trigonometric Series

Let F(z) be an analytic (differentiable) function of the complex variable z = x + iy, without sing~larities2~tbr lzl < 1. Then, it is well known that

2 A point z is said to be a singularity of the function F(z) if in the complex plane there exists no circle with center at z (no matter how small its radius) within which F(z) is analytic.

Page 120: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

106 TRIQONOMMTRIC SBR~BS WITH DECREASING COEFFICIENTS CHAP. 4

F(z) can be expanded in a power series

for 151 1, Leo, within the circle of radius 1 with center at the origin of the complex plane. Suppose that the coefficients of this series are real n u m b , and set z = eix. Then (5.1) still holds, since

Thus, for any x, we have

F(e") = co + c1& + c2C(Z* + + cndW + = co + cl(cosx + isinx) + c2(c<w2x + 1sin2x) + @ -

+ c,(cosnx + isinax) + - - a (5-2) = (co + c1c0sx + c t ~ 2 x + * * * + cncosm + * * * )

+ I(cl sin x + c2 sin 2* + + Cn sin nx + 0 ) .

We now separate the real and imaginary parts of F(&), i.e., we write F(erx) in the form

F(c") = f(x) + ig(x), where f(x) and g(x) are real functions. Then, it obviously follows from (5.2) that

f ( ~ ) = c0 + c ~ C O S X + c2WS2* + a * * + cnCOQnX + * * * ,

g(x) = cl sinx + c2sin2x + - 9 - + c,sinnx + a * = .

This fact can be used to find the sums of certain trigonometric series.

Example 1. It is well known that

for all z. Then, by (5.2) we have

cos 2x cos nx n!

( sin 2x sin nx + i sinx + - + - - - + + . . .).

2! n ! But

= e x + l r inx = p x e i r i n x

= em" [cos (sin x) + i sin (sin x)], and therefore

cos 2x cos nx e ~ a x cos (sin x) = 1 + cos x + + * * * + +...

2! n! 9

sin 2x sin nx eco8xsin(sinx) = sinx +- + a - + - + - 0 . 2! n!

Page 121: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

In this example, we started from a given function of a complex variable and found that its real and imaginary parts can be expanded in trigonometric series. In other words, we have found a new approach (as compared with Chs. 1 and 3) to the problem of expanding a function in trigonometric series. Moreover, in many cases, the argument given at the beginning of this section also allows us to solve the inverse problem, i.e., to find the sum of a given trigonometric series. In fact, suppose we are given two convergent series

Using these series, we form the series

(co + qcosx + c2cos2x + - 0 . ) + i(cl sinx + qsin2x + - - 0 )

= q + cl(cos x + i sin x) + c2(cos 2x + i sin 2x) +. . = co + el& + c2ei& +*-,

with complex terms. Denoting e" by z, we obtain the power series

If we know the sum F(z) of this series for the values of t of interest, then, setting

F(eix) = f (x) + ig(x),

we obviously have

f (x) = co + Cl cosx + qcos2x + * * * ,

g(x) = el sinx + q d n 2 x + * m e .

Example 2. Find the sums of the series

cosx cos2x + cos nx 1 + - + - * a + +..*

P p2 P 9

sin x sin 2x - sin nx +. . . (5.3) + p2

+ - - * + P P" 9

where p is a real constant with absolute value greater than 1. The two series (5.3) converge for all r Consider the series

+ sin x sin 2x +...) + i(- + - +... P p2

We have

Page 122: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

108 TRIWNOME~[~\IC SER~BS w r r ~ DECREASING ~ ~ ~ I C I E N T S CHAP. 4

since the series on the left is a geometric series, which converges for lz/pl < 1. Therefore

F(etx) = p - - P p - eix @ - cos x) - i sin x

- - (p - cosx) + isinx ( p - cos x) + i sin x = P p2-Zpcosx+1 ?

(p - cos x)2 + sin2 x

and we find that

p(p - cos x) cos x cos 2x cos nx + * * * + +..* I = I + - p2-2pcosx+ 1 P p2 P +-

p a n ~ sin x sin 2x =- + sin nx +.*.+- +... p2-*cosx+ 1 P p2 P"

for all x.

6. A Stronger Form of the Results of Sec 5

At the beginning of Sec. 5, we assumed that the function F(z) had no singularities for lzI 6 1. Suppose now that F(z) has no singularities only for lzl < 1, while F(s) has both rgulm points3 and singularities for lzI = 1 (in geometrical language, on the unit circle C with center at the origin). Then, for 121 < 1, the expansion (5.1) is still valid. Moreover, (5.1) Ir still valid at every regular point z on the circle C at which the series (5.1) converges. We now prove this result, for the benefit of the reader who may not already know it. First we need the following

LEMMA. Let

be a convergent series (with real or complex terms). Then the series

converges for 0 ( r ( 1, and its sum 4r) C continuous on the interval 10, 11.

Proof. If

then for every E > 0, there exists an integer N such that if n 2 N

3 I.e., points which are not singularities. (~oltslatorr)

Page 123: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Set. 6 TRIQONOMETIUC SERIES WITH DECREASING COEFFICIENTS 1 09

where an is the partial sum of the series (6.1). Consider the remainder of the series (6.1)

and its partial sums

&+I + U-2 + + un+m (m = 1,2,. . .). By (6.3) we obviously have

Thus, every partial sum of the series (6.4), just like its entire sum, is ( r in absolute value. Since the numbers

dwrease monotonically to zero as n + co for any r (0 ( r < I), Abel's lemma is applicable to the series

It follows that this series, and hence the series (6.2) converges, and moreover that

Now, letting an(r) denote the partial sum of the series (6.2), we have

for 0 ( r < 1. Noting that 41) = a and 4 1 ) = an, and using (6.3). we can assert that the inequality (6.5) holds everywhere on the interval 0 ( r < 1. This proves that the series (6.2) converges uniformly on 0 < r < 1 and hence implies that the function a(r) is continuous on 0 < r 4 1. This proves the lemma.

To prove the proposition formulated at the beginning of this section, we assume that the series

converges at a point z lying on the circle C. By the lemma just proved, the function

is a continuous function of r for 0 6 r < 1, and therefore

lim F(rz) = lim (co + clrz + czr2z2 + 9 ) r-+I r+l

r< 1 r<l (6.6)

Page 124: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

1 10 T R I a O N O ~ C SERIES WITH D-0 COBFFICIBNTS CHAP. 4

On the other hand, since the point z is a regular point, F(z) is continuous at z, and therefore

since rz + z as r + 1. Comparing (6.6) and (6.n we obtain

as was to be shown.

Thia result makes the argument of Sec. 5 legitimate for any regular point z = dx (such a point lies on C, since l&x( = 1) for which the series in (5.1) converges. To illustrate the situation, we now consider two examples.

Example 1. It is well known that

for 121 < 1, and that the function in (1 + z) is analytic at all points of the unit circle C except the point z = - 1, or equivalently, the point z = d(u+l)n. By (5.2), if z = Crx where x # (2k + l)rr, we-have

where we are justified in writing the equation, since the two series on the right actually converge for x # (2 + 1)rs. (See Theorem 1 of Sec. 4.) On the other hand, for - x i x < x, we obviously have

1 + &x= (1 + cos x) + isin x

for -x < x < x. Then, by (6.8) we find that

ln(2oos:) = msx - cos 2x cos 3x + 3

-... 2 9

- - X - = sin x - sin 2x sin 3x

+ 3 - 6 . .

2 2

4 We have used the following familiar property of the logarithm: If z = geie, -n < 8 < rc, then In z = In p + lo. In our case p = 2 cos (x/2), 8 = x/2.

Page 125: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 6 TRIGONOMETRIC SERIES WITH DECREASING COEFFICIENTS I 1 1

for -rr < x < z, and we have obtained two expansions with which we are already familiar [see Example 2 of Ch. 3, Sec. 14 and formula (13.9) of Ch. 1, Sec. 131.

By a similar method, we can find the two expansions cos 2x + cos 3x

2 3 + a * - ,

(0 < x < 2%) (6.10) X - x sin 2x sin 3x

= sin x + + 3

+... 2 2 9

which we have also encountered previously. In this case, we have to start from the function

for which X - X f(x)= -h g(x)= (0 < x < %).

However, a much simpler way of obtaining the expansions (6.10) is to make the substitution x = t - x in (6.9).

Example 2. Find the sums of the series

cos 2x cos 3x cos (n + 1)x + 2.3 + * " + n(n + 1)

+... 1 a2 9

sin 2x sin 3x sin (n + 1)x 1-2 + 2.3 + * " + n(n + 1)

+ * * a .

These series converge for all r Consider the new series

cos2x cos3x sin 2x sin 3x ( 1*2 + 2*3 + ) + i 1.2 + 2.3 + - - - )

It follows from the identity 1 1 1

n(n+ i n + 1 that

Page 126: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

I 12 TRIOONOMETRIC SERIES WITH DECREASINO COEFPICIBNTS CHAP. 4

(see Example 1) for all z satisflying the conditions is1 ( 1, z # - 1. There- fore,

F(dx) = (1 - err) ln (1 - e") + &x

= [(I - cos x) - i sin x]

+ (cos x + I sin x)

= [(I - w s x) 1n 2 sin - - ( 3 "iX sin x + ms x]

- X + I [T (cor x - 1) - sin x ln

For 0 < x < 2rr (see Enample l), and consequently c o s b cos 3x %inx + cosx = (1 -cosx)ln +- +... 2-3 9

X - X - (00s x - 1) - sin x ln sin 2x sin 3x 2 1.2 + - + * * * . 2.3

PROBLEMS 1. For which values of x do the following series converge:

00 sinnx c o s n ; ~ + shnx ?

n= 1

2. For which values of x are the sums of the series in Prob. 1 continuous? Are these series the Fourier series of square integrable functions?

3. For which values of x do the following series converge: aD

00s nx cos nr + (-ly sin nxa Inn 9

n o 1 n=2

4. For which values of x are the sums of the series in Prob. 3 continuous? Construct the graph of the sum of the series c).

5. For the complex variable z = x + iy, the trigonometric and hyperbolic functions are defined by the series

23 2s sinhz=z+-+--+*-- 31 S! (hyperbolic sine),

22 z(o coshz= 1 + - + - + * * * 21 41 (hyperbolic cosine).

Page 127: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

PROBLEMS TRIOONOMETRIC SERneS WITH DECREASINO COEFPIClBN7S 1 1 3

Using the formulas sin(or + p) = sinacosg + cosasinp, cos(a + p) = cosacosg - sinorsing,

which are also valid for complex or and g, prove that sin (a + @) = sin a a s h p + f cos a sinh A cos(a + @) = cosacmshg - isinasinh(3.

6. Find the sums of the series

sin3x sinsx b) sin x - - +-- ... 31 5!

Hint. F(z) = sin z (see Sec. 5). Use the first of the formulas (I) of Prob. 5.

7. Find the sums of the series

sin2x sin4x sin& b) --- + -- ... 21: 4! 61

Hint. F(z) = cos z. Use the second of the formulas (I) of Prob. 5.

8, Find the sums of the series

sin x sin 2x sin 3x b) ---+ --.... 1.2 2.3 3 *4

Hint. Use the method of Example 2 of Sec. 6 and the result of Example 1 of the same section.

9. Find the sums of the series

sin 2x sin 3x b) - 0 - + . . .

sin nx 3 8

+ (-lYi +"'. n* - Hint. Use the identity

and the results of Example 1 of Sec. 6.

10. Find the sums of the series

a) 2cos2x 3cos3x ncosnx

3 - 8 (-1)" n, - 1 + . a o ,

b) 2sin2x 3sin3x nsinnx

3 - 8 + * * * + (-lPqn - + * * a

Page 128: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

CHAP. 4

Hint. Use the identity

and the results of Example 1 of Sec. 6.

11. Find the sums of the series

sinx sin% sin nx +- b, I+p 2 + p +...+ - +...

n + p 9

where p is a positive integer.

Hint.

Also use the results of Example 1 of Sec. 6.

12. Show that if 2 an converges, then so does 2 50 n u 1 n - 1 n

13. Using the notation of Sec. 1, show that if la,J < M, if a,, + 0 as n + a, and if

00

then 2 anun converges.

14. Show that if lat[ s 141 3 --+O and if

converges absolutely at even a single point xo, then

Show that the same is true in the case of the series

provided that xo is not an integral multiple of n.

Page 129: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

OPERATIONS ON FOURIER SERIES

I. Approximation of Functions by Trigonometric Polynomials

In Ch. 3, we established certain conditions under which a continuous (and sometimes even a discontinuous) function of period 2n can be repre sented as a sum of a trigonometric series. This raises the following question: Was it only because of the inadequacy of the argument used that we were not able to prove that the Fourier series o! any continuous function f(x) always converges to f(x)? It turns out that the answer to this question is negative, and in fact there exist examples of continuous functions with discontinuous Fourier series.

We now adopt a different approach, i.e., we consider the approximate representation of functions by trigonometric series. Here we immediately obtain the following remarkable result.

THEOREM. Let f(x) be a continuous fictiofi of period 2rr. Then, given any e z 0, there exists a trigonometric polynomial

for which

for any X. I I S

Page 130: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

1 1 6 OPERATIONS ON FOURIER SERIES CHAP. 5

Proof. Consider the graph of the function y = f(x) on the interval [-z, x]. Divide [-x, x] into subintervals by the points

and construct the continuous function g(x) for which Axk) = f(xk) (k = 0, 1,2,. . ., m) and which is linear on each subinterval [x&-~, xk]. The graph of the function y = g(x) is a broken line with vertices on

the curve y = f(x) (see Fig. 40). The subintervals into which we divided [ox, x] can be made so small that the inequality

is satisfied for any x in the interval [-IC, n]. Moreover, the function g(x) can be periodically extended onto the whole x-axis. Then, ob- viously, the condition (1.2) is also valid for the periodic extension of g(x), which is continuous and piecewise smooth on the whole x-axis.

We now form the Fourier series of g(x). According to Theorem 2 of Ch. 3, Sec. 10, this series converges uniformly to g(x). But then for all sufficiently large n we have

for any x, where o,,(x) denotes the nth partial sum of the Fourier series of the function g(x). Let n be an index for which (1.3) is valid. Then it follows from (1.2) and (1.3) that

for any x, which proves the theorem.

We now note the following consequences of this theorem:

Page 131: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 1 OPERATIONS ON FOURIER SERIES 1 17

COROLLARY 1. f(x) is continuous on the interval [a, a + 2x1 and f(a) = f(a + 24 , then for any E > 0, there exists a trigonometric polynomial of the form (1. 1) for which

for any x in the interval [a, a + 2x1.

To see this, it suffices to make the periodic extension of f(x) onto the whole x-axis (because of the condition f(a) = f(a + 2x), this does not destroy the continuity) and then apply the theorem to the extended function.

COROLLARY 2. If f(x) )b continuow on an intend [a, b] of length less than 2x, then for any c z 0, there exMs a trigonometric polynomial of the form (1.1) for which

If(x) - on(x)I &

for any x in [a, b].

Roo$ Extend f(x) from [a, b] onto [a, a + 2x1 in such a way that continuity is preserved and the extended function satisfies f(a) = f (a + 279. For example, this can be done by setting f(a + h) = f (a) and then taking the function to be linear on the interval [b, a + 2z] (see Fig. 41). Then Corollary 1 is applicable to the extension off (x)

obtained in this way, and therefore the inequality (1.4) is valid for all x in [a, a + k], in particular, to all x in [a, b].

What has just been proved can be formulated succinctly as follows: Under the conditions of the theorem (or its corollaries), the fwrction f(x) can be un$iorrnly approximated by a trigonometric polpomlal of the form (1.1) with any degree of accuracy specffied in advance.

2. Completeness of the Trigonometric System

In view of Ch. 2, Sec. 9 (the completeness criterion for an orthogonal system), to prove that the trigonometric system is complete it is sufficient to

Page 132: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

1 1 8 OPERATIONS ON FOURIER SERIES CHAP. 5

prove that given any function f(x) which is continuous on [-z, n] and any E > 0, there exists a trigonometric polynomial an(%) for which

We now prove that this is actually the case. If f(-x) = f(rr), then by Corollary 1 of the theorem of Sec. 1, there

exists a trigonometric polynomial an(x) for which

If (x) - an(x)l * - for any r Therefore, we have

as required. Next, suppose that f(-rr) # f(rr). Let M be the maximum of the

absolute value of f(x) on the interval [-rr, x], and choose the number h > 0 so small that the condition

holds. Now, let g(x) be the continuous function which coincides with f(x) on the interval [-z, z - h], is equal to f(-z) at x = z, and is linear on the interval [rr - h, rr]. (This construction is like the one shown in Fig. 41.) Obviously, lg(x)l < M, and hence

On the other hand, g(x) is continuous and takes equal values at the end points of the interval [-z, x]. Therefore, then exists a trigonometric polynomial %(x) for which

But then it follows from (2.2), (2.3) and the elementary inequality

(a + b)2 < 2(a2 + b2) that

as was to be proved.

Page 133: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 3 OPERATIONS ON FOURIER SERIES 1 19

3. Paneval's Theorem. The Most l mportant Consequences of the Completeness of the Trigonometric System

Since the trigonometric system is complete, Bessel's inequality [see (1.2) of Ch. 31 becomes the equality

where f(x) is an arbitrary square integrable function, and a , bn are its Fourier coefficients. Formula (3.1) is known as Parseval's theorem.1 Recalling Theorem 1 of Ch. 2, Sec. 7 and the fact that

we obtain THEOREM 1. If f(x) and F(x) are square integrable functions defined

on [-x, x], for which a0

+ 2 (an cos nx + bn sin nx), f(x) - 7 n-1

00

F(x) - 2 + 2 (A,, cos nx + B,, sin nx), n = l

The following result is an immediate consequence of Theorem 2 of Ch. 2, Sec. 7:

THEOREM 2. The trigonometric Fourier series of any square integrable function f(x) converges to f(x) in the mean, i.e.,

n

lim I" [/(XI - (? + 2 (ak a)s kx + bk sin kx) tii = 0. -00 -It k= I )I2

This theorem is all the more remarkable, in view of the fact already noted in Sec. 1 that a trigonometric Fourier series does not always converge in the ordinary sense even for a continuous function.

In Ch. 2, Sec. 7, we proved that a Fourier series can converge in the mean

1 Attributed by the author to the Russian mathematician A. M. Lyapunov "who first g;ave a rigorous proof of the formula (for the case of bounded functions)." Theorem 1 below, of which (3.1) is a special case, is the result usually called Parseual's theorem in the English-language literature. (Tramlator)

Page 134: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

120 OPERATIONS ON FOURIER SERIBS CHAP. 5

to only one function f(x) [to within a change of f(x) at a finite number of points*]. This implies

THEOREM 3. Any square integrable fmction f(x) is completely defined (except for its values at a finite number of points) by its trigonometric Fourier series, whether or not the series converges.

In Theorem 3 we say "defined" but this does not mean that we yet know how to actually determine a function from a knowledge of its Fourier series. The problem of actually "reconstructing" a function from its Fourier series will be solved completely in the next chapter; however, in some special cases, the problem can be solved by using the method given in Ch. 4, Secs. 5 and 6.

Finally, we note two more consequences of the completeness of the trigonometric system (see Theorems 1 and 2 of Ch. 2, Sec. 8) :

THEOREM 4. Any continuous function f (x) which Is orthogonal to all the fwrctionr of the trigonometric system must be identically zero.

In other words, except for the trivial function which is identically zero, we cannot add an extra continuous function to the trigonometric system and still have an orthogonal "enlarged" system. Theorem 4 can also be paraphrased as follows: If all the Fourier coeflcients of a continuousfMelion are zero, then the fmction vanishes identically.

THEOREM 5. If the trigonometric Fourier series of a continuous fwction f(x) i( unformly convergent, then the sum of the series equals f(x).

*4. Approximation of Functions by Polynomials

As a rather simple consequence of the results of Sec. 1, we can obtain the following result which is often very useful :

TwD~esr.3 Let f(x) be a fwction which tr continuous on the interval [a, b]. ntm, for any e > 0, there exbts a polynomial

pm(x) = co + C I X + *cmP,

for which

everywhere on the interval [a, 61.

2 See the footnote to Ch. 2, Sec. 7. (Trunslator) 3This proposition is usually known as the Weierstrass approximation theorem.

th an slat or)

Page 135: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

OPERATIONS ON FOURIER SERIES I 2 I

Prwf. By making the transformation

or equivalently

we take the interval [a, b] of the x-axis into the interval [O,N of the z-axis. (If the length of [a, b] is less than Z~C, we do not have to make this transformation.) We now set

b - a f (- 7C t + a) = ~ ( t ) .

By Corollary 2 of the theorem of Sec. 1, there exists a trigonometric polynomial

n

o,,(t) = a0 + 2 (ak cos kt + & sin kt), k= l

for which

everywhere on [O, x]. Holding n fixed, we choose a positive number u > 0 which is so small that the condition

Now, it is known that for all z

where these series converge uniformly on every interval of finite length, in particular, on 0 < z $ nrr. But then, for fixed n and all sufficiently larg ' we have

Page 136: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

122 OPERATIONS ON FOURIER SBRIES CHAP. 5

where k = 1,2,. . ., n, for any t in the interval [0, rr]. (Since there are only a finite number of the functions cos kt, sin kt for k = 1, 2,. . ., n, ir., just Z of them, the index 1 can be chosen large enough to make all the inequalities (4.5) hold simultmeo11~ly.)

We denote the polynomials of degree 21 and 2l+ 1 appearing in parentheses in the inequalities (4.5) by rk(t) and sk(t), respectively. Then

lcos kt - rdt) 1 C a, 1 sin kt - sk(t)l < 0

(k = 1,2,. . ., n) for any t in [O,d. Consider the sum

which is a polynomial of degree rn = 2l + 1. By (4.4) and (44, this polynomial satisfies the inequality

everywhere on [0, rr]. Therefore, it follows from (4.3) that

everywhere on [0, x]. Using the formula (4.2) to return to the variable q, we obtain

eveqwhere on [a, b]. It is easy to see that the function

is also a polynomial of degree m. Consequently, (4.7) is just the inequality (4.1) needed for the proof.

5. Addition and Subtraction of Fourier Series. Multiplication of a Fourier Series by a Number

To obtain the Fourier series of the sum or difference of two hrnctions whose Fourier series are known, it is sufficient to add or subtract the two

Page 137: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

known series. In fact, suppose that 00

f(x) - 3 + 2 (a, cos nx + b, sin nx), nu1

aD

F(x) - -?-O + 2 (A, cos nx + B, sin nx). 2

Then, the Fourier coefficients an and B, of the function f(x) f F(x) are given by

1 . = - 1 x I-, f(x) cos nx dx f - In F(x) sin nx dx = an f A,

x -n

and similarly

which proves our assertion. It can be shown in just the same way that the Fourier series of the function

kf(x), where k is a constant, is obtained from the Fourier series of f(x) by multiplying all its terms by k.

Despite the simplicity of these considerations, they testify to a very important fact, namely, that although convergence is not assumed, it is possible to operate on Fourier series as if they were convergent, i.e., as if the sign = appeared instead of -. The same sort of phenomenon will also be encountered in subsequent sections of this chapter.

*6. Products of Fourier Series

How do we form the Fourier series of the product f (x)F(x), if we know the Fourier series of the factors? To answer this question, we argue as follows: We begin by assuming that f(x) and F(x) are square integrable functions, in which case we know that the product f(x)F(x) is an integrable function (see Ch. 2, Sec. 4). [Note that if we fail to impose this requirement on f(x) and F(x), the product f(x)F(x) may turn out to be nonintegrable, and then the problem of finding the Fourier series of f(x)F(x) becomes meaning- less.]

Now, suppose that f(x) and F(x) obey the relations (5.1) and that

ao + 2 (a,, cos nx + B, sin nx). f(w'(x) - 7 n= 1

Page 138: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

124 OPERATIONS ON PO- S E R ~ ~ S CHAP. 5

Our problem is to express the coefficients a, and in terms of a,,, bn, A,, and Bn. Using Theorem 1 of Sec. 3, we find that

1 a do QD

ao = x - Lscf(x)F(x) dk = + 2 (a,,A,, + bnBJa n = l

To calculate

it is sufficient to know the Fourier coefficients of the function F(x) cos nx, since then we can use Theorem 1 of Sec. 3 again, this time applied to the product f (x)F(x) cos n r The Fourier coefficients of F(x) cos nx are

1 ; j'=% F(x) cos nx cos nuc dx

which becomes

+(Am+,, + Am-J for m b n, +(A,, + A,,,,) for. m < n.

If we agree to write A,& = A&,

then we have

Similarly, we find

1 - Ln F(x)cosnxsinmxdx = ? It

' [A rsc ~ ( x ) sin (m + n)x tix

1 + - r F(x) sin (m - n)x dx] It -n

where we have set

Thus, we now have the Fourier coefficients of the function F(x) cos nx. Therefore, applying Theorem 1 of Sec. 3 to the integral (6.2), we obtain

Page 139: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

In just the same way, we find that

The formulas (6.1), (6.3), and (6.4) give the solution of our problem. We note that these formulas can also be obtained by formally multiplying together the two series (5.1) (ic., by acting as if these series are convergent and multiplication is justified4), then replacing products of sines and cosines by their sums and differences, and finally collecting similar terms.

7. Integration of Fourier Series

In the applications, one encounters cases where only the Fourier series of a function is known, but not the function - itself. In this connection, the following problems arise :

1) Given the Fourier series of the function f(x) of period 2rr, calculate

where [a, b] is an arbitrary intend; 2) Given the Fourier series of the function f(x), find the Fourier series of

the function

F(x) = f (x) dx. 0

The solution to the first problem is given by

THEORFM 1. If the absolutely integrable fmction f(x) of period 2rr is specNed by giving its Fourier series

00

f(x) + 2 (an cos nx + bn sin nx), n = l

then

4 It is wellknown that the product of two convergent series 8 = ul + u2 + + ua + andg = vl + 112 + - - a + & is given by the formula

which is vdnlid whenever the last series converges. In particular, the formula dwqyr holds if the two series converge absolutely.

Page 140: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

126 OPERATIONS ON POWBR S E R ~ CHAP. 5

can be found by term by term integration of the series (7.1), whether or not the series converges, toe.,

+ 5 an(sin nb - sin na) - bn(ms nh - cos M) n

(7.2) n o 1

Proof. If f(x) is square integrable, then this theorem is a special case of Theorem 3 of Ch. 2, Sec. 8. However, in the general case, we argue as follows: First we set

This function is continuous and has ar. absolutely htegrable der' ,ative (which possibly does not exist at a fiuite number 'PC:- Moreover

so that F(x) has period 2x. Therefore, F(x) can be expanded in Fourier series (see Ch. 3, Sec. 1 1) :

PO

+ 2 (An cos ~ Z X + B, sin nx). F(x) = n=1

Integrating by parts for n 3 1 gives

sin nx X''F = x !. [ . x ) ~ ] XP-z - z n 113, [f (~) - 21 sinnxtix = - -, bn n

and similarly

Therefore 00

F(x) = $ + 2 a,, sin PIX - bn cos nx n = l n

Page 141: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

a @ + $ + e an sin nx - bm cos nx f(x) dx = o n = l n (7-4)

To obtain (7.2), it remains only to set x = b, then x = a and subtract the resulting formulas.

The solution to the second problem is given by

THEOREM 2. Let the absolutely integrable function f(x) be spec@ed by giving its Fourier series

f ( ~ ) .Y + 2 (an cos nx + b,, sin nx), n= 1

whether convergent or not. Then the integral of f(x) luu the Fourier expansion

Proof. We set x = 0 in (7.4), thereby obtaining

On the other hand, by formula (13.9) of Ch. 1, we have

QJ x sin nx - = 2 (-ly+l- 2 n = l n

for - x < x c x. Substituting (7.6) and (7.7) in (7.4) gives (7.5).

Remark. In passing, we have proved that the series

converges for any absolutely integrable function. This fact is some- times useful, since in some cases it allows us to distinguish the Fourier series of absolutely integrable functions from other trigonometric series. For example, the series

Page 142: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

128 OPERATIONS ON FOURIER SBRIBS CHAP. 5

which converges everywhere (see Sec. 3 of Ch. 4) cannot be the Fourier series of an absolutely integrable function, since the series

diverges. Next, we note an important special case of Theorem 2.

~ R B M 3. If a. = 0 and the other conditions are the same as in Theorem 2, then

- bn cos nx + a,, sin nx 9

n = I n = l n (7.8)

for all x,s i.e., the Fourier series of the integral can be obtairred by term by term integration of the series for f(x).

ProoJ The formula (7.8) is obtained from (7.5) by simply setting a0 = 0. Its validity for all x [and not just for -X < x < z, as in (7.5)] follows from the periodicity of the integral on the left in (7.8), which itself follows from

We can use (7.8) to obtain many new Fourier series, by starting from a known series. For example, we know that

x sin 2x sin 3x - = sin%- - + 3

-... 2 2 (-X < X < x).

Integrating, we obtain

cos 2x cos 3x - (msx - + 32 0 . . .) 22 a0

= C - 2 (- 1y+l cos nx n2

C = const. n = l .

To find C, we integrate the last equation from -x to x. Since the series on

5 If we are interested only in the values -x < x < z, then as in Theorem 2, we do not have to require that f(x) be periodic. However, if we are iaterested in all values of x, then for (7.8) to be valid, we have to assume that f (x) is periodic.

Page 143: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

the right converges uniformly, the term by tenn integration is justified, and we obtain

Therefore, we have

so that

an expression that we have already obtained in Ch. 1, Sec. 13. Integrating this formula again gives

x x2 sin nx ( 0 E D - f ) d x = 2 ( - l y + ' - n=l n3

a0

- 0 s sin nx x2x x3 Z(-ly+l-*

12 12 ,#=I n3

8. Differentiation of Fourier Series. The Case of a Continuous Function of Period 2%

THEOREM 1. Let f(x) be a continuous fmction of period 2x, with an absolutely integrable derivative (which may not exist at certain points9. Tihen the Fourier series of f ( x ) can be obtaincd from the Fourier series of the fmction f (x) by term by term diferentiation.

Proof. Let a0

f ( ~ ) = + 2 (an cos nx + bn sin nx), n=l

where we can write the equality because of the theorem of Ch. 3, Sec. 11. Let a; and b; denote the Fourier coefficients off(x). First of all, we have

6 In other words, f'(x) may not exist at a finite number of points (in each period).

Page 144: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

CHAP. 5

Then, integrating by parts, we obtain

-It n = [ f ) 0 s ] + ; I f(x) sin nx dx = nk,

'IC --It -IC I - (8.21 . . : 1 f'(x) sin nx dw C - -,

f (x) cos nx dx = -nu,,.

Therefore a0

f ' ( ~ ) - 2 n(bn cos n~ - an sin nx), n o 1

and this is the series obtained from (8.1) by term by term differentiation.

Remark. Under the conditions of Theorem 1

b: a: a,, = - -9 bn = n n

which follows at once from (8.2). Moreover, since the Fourier co- efscients of an absolutely integrable function converge to zero as n + a, (see Ch. 3, Sec. 2), we can write

lirn nu,, = lim nbn = 0, *a0

i.e., a, and bn go to zero faster than lln as n + 00.

2 Let f (x) be a continuous fmctbn of period 27t, which has m derivatives, where m - 1 derhratives are continuous and the m'th derhttve is absolutely integrable (the m'th derivative may not exht at certain points). TIten, the Fourier series of all m derivatiues can be obtained by term by term diferentiation of the Fourier series of f(x), where all the serh, except posszbly the last, converge to the corresponding derivatives. Moreover, the Fourler coeflcients of the fmction f(x) satrrfv the relations

lim nmam = lirn pbn = 0. -OD -00

ProoJ To prove the first assertion, it is sufficient to make m appli- cations of Theorem 1. The convergence of all the series obtained by term by term differentiation, except possibly the last, to the correspond- ing derivatives follows from the differentiability of these derivatives [up to the (m - 1)th derivative].

Page 145: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SBC. 8 OPERATIONS ON FOURIER SERIES 1 3 1

The relations (8.4) are obtained as a result of successively applying the relations (8.3) m times. Thus

where a:, 4 . . . , b:, b:, . . . are the Fourier coefficients of the functions f'(x), f "(x), . . . , while a,, and B, denote the appropriate Fourier cb efficients of the function f(ll)(x), taken with the proper sign. Since f (mxx) is absolutely integrable, a,, + 0 and B,, + 0 as n -t a, which implies (8.4).

Remark. Under the conditions of Theorem 2, the series for f(x) and all the series obtained from f(x) by term by term differentiation, except possibly for the last, converge uniformly (this follows from Ch* 3, Sec* 11).

The next proposition is in a certain sense the converse of Theorem 2:

THEOREM 3. Given the trigonometric series a0

00 + 2 (an cos nx + bn sin nx), * n= l

if the relations

me satlsjFed by the coeBcients a,, and bn, then the sum of the series (8.6) is a continuous fwtction of period 2z, with m - 2 continuous derivatives, which can be obtained by term by teim diflerentiation of the series (8.6).

Proof. Denote the sum of the series (8.6) by f (x). By (8.7), we can write

f(x) = 3 + 2 (* nm cos nx + nm sin nx),

where

If we formally differentiate this series, then after k differentiations the coefficients have absolute values which do not exceed the numbers

Page 146: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

132 OPERAT~ONS ON ~ 0 - - CHAP. 5

It follows that the sum of the absolute values of the coefficients con- verges for k = 1,2,. . ., m - 2. Therefore, by Theorem 1 of ch. 3, Sec. 10, the series obtained by term by term differentiation of the series (8.6) are uniformly convergent for k = 1.2, . . . , m - 2. But then, it follows from Theorem 2 of Ch. 1, Sec. 4 that the function f(x) is differentiable m - 2 times (and hence continuous), that the deriva- tives are continuous, and that the term by term differentiation is legitimate.

*9. Differentiation of Fourier Series. The Case of a Function Defined on the Interval [ - X , x ]

THEOREM 1. Let f(x) be a continnuow function defined on the intend [ - X , n] with an absolutely integrable derivative (which may not exist at certain points). llbrn

C 09

S(X) - 2 + 2 [(nb,, + (- 1 ) " ~ ) ws nx - na,, sin nx], (9.1) n o 1

w k e an and bn are the Fourier coeflcients of the fmct io~ f(x) and the constant c is given by the formula

Proof. Let

ab a0

f (x) - T + 2 (a: cos rur + b: sin nx), n o 1

so that

Obviously we have7

4 QO

f (x) - -i. - 2 (a: cos nx + b: sin nx). nu1

The Fourier series of the function

7 Although (9.3) is not an equality, it is still possible to transpose tenns from the right- hand side to the left-hand side, as can be verified by calculating the Fourier cafficients of the fmction appearing in the left-hand side. In our case, the constant term of the function f'(x) - hb vanishes, and all the other coefficients remain the same as for f'(x).

Page 147: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 9 OPERATIONS ON FOURIER SERIES 1 33

can be obtained from the series (9.3) by term by term integration (see Theorem 3 of Sec. 7). ~hereiore, conversely, the series (9.3) can be obtained by term by term differentiation of the series for the function (9.4). But we have

PO

f(x) = 2 + 2 (an cos nx + bn sin nx), n o 1

and by (7.7)

Therefore

a' QD

f (x) - - - 2 [-nan sin nx + (nb,, + (- 1)"~') cos nx], 2 n= 1

which implies (9.1). if we set c = 4.

COROLLARY. If c = 0, i. e., f (x) = f( - x), then (9.1) gives

QO

f (x) - 2 n(bn cos nx - an sin nx). n = l

In other words, if f(n) =A-IC), the Fourier series of f(x) can be differentiated term by term. But this is immediately clear, since if f(x) = f(-n), the periodic extension of f(x) onto the whole x-axis is continuous and hence Theorem 1 of Sec. 8 can be applied.

Remark. Theorem 1 is especially important in the case where the Fourier series of f(x) is given instead of f(x) itself, for it shows that a knowledge of the Fourier series of f(x) is all that is needed to form the Fourier series of f'(x). However, in this case, the calculation of the constant c can turn out to be difficult if we use the formula (9.2). We can avoid the use of (9.2) by recalling that the Fourier coefficients of an absolutely integrable function converge to zero as n + a, (see Ch. 3, Sec. 2). Therefore, it follows from (9.1) that

lim [&,, + (- lyc] = 0, n+PO

whence

It is usually an easy matter to calculate this limit. Moreover, it is not

Page 148: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

134 mmunom ON ~ 0 - ~BRIES CHAP. 5

hard to see that the existence of this limit is equivalent to the quantity nbn having limits for even and odd n separatelyy which are equal in absolute value and opposite in sign. Theorem 1 presupposes that we know that the function f(x) is continuous

on [-zy n] and that f(x) has an absolutely integrable derivative. In the applications, one often encounters situations when we only know the Fourier series of f(x). In this case, the problem becomes more complicated, i.e., from a knowledge of the Fourier series we have to ascertain whether f(x) is differentiable and whether its derivative is absolutely integrable, and if the answer is aflirmative, we have to form the Fourier series of the derivative. The following theorem often helps to solve this problem:

THEOREM 2. Suppose that we are given the series 00

9 + 2 (a, cos nx + bn sin nx). * n=l

the series

C 00

+ 2 [(nb,, + (- 1)"c) cos nx - na,, sin 12x1, 2 n=l

where c = lim [(- 1 ~ + 1 ~ ] ,

n4a,

is the Fourier series of an absolutely integrable function q(x),8 tho, the series (9.5) is the Fourier series of the function

which is continuous for -z < x < n. Moreoaer, (9.5) converges to f(x), and obviousk'yf'(x) = cp(x) at all the continuity points 4cp(x).

Proof. We can apply Theorem 2 of Sec. 7 to the series C

00

q ( ~ ) - 3 + 2 [(nbn + (- I ) % ) ws nx - na,, sin nx] n= 1

obtaining

00 na,, cos nx + [nbn + (- 1)"c + (- l)"+'c] sin nx = - 5 ' n - + 5'

= - 2 an + 2 (a,, cos nx + b,, sin nx) n o 1 n=1

8 It is not assumed that the series (9.6) converges.

Page 149: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 9 OPERATIONS ON FOURIER SERIES 1 35

for - x < x < x. Thus, we have

00 00

C T(X) d~ + 2 an = 2 (an cos ~ I X + bn sin nx) n u 1 n= 1

and hence

a0 a0 Ji T(X) d~ + + 2 an = 9 + 2 (a,, cos nx + bn sin M). n=l n = l

Example 1. The series 00 n sin nx 2 (-I)" ns - 1

n=2

is the Fourier series of a continuous9 and differentiable function for -x < x < x. To see this, we first use (9.7) to find

and then form the series (9.6). The result is

1 00

+ cos x + 2 [(- lr n2

- 2 n2 - 1 + (- l)"+l] cos nx n=2

1 00

- - + c o s x + C ( - l y COS ItX

2 n = 2 n2 - 1'

This series converges absolutely and uniformly (since the sum of the absolute values of the coefficients is obviously convergent), and therefore it has a continuous sum p(x) of which it is the Fourier series (see Theorem 1 of Ch. 1, Sec. 6). According to Theorem 2

00 n sin nx n = 2

and 00

f'(x) = cp(x) = - 2 + cosx + 2 (-ly cos nx n = 2 n2 - 1'

Incidentally, we note that it is sometimes possible to find the sum of

9 The continuity of the sum of the series can be inferred from Theorem 1 of Ch. 4, Sec. 4.

Page 150: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

136 OPBRATIONS ON FOURIER smm CHAP. 5

a Fourier series by differentiating it. Thus, in this example, if we apply Theorem 2 to the series (9.9), we obtain

00 n sin nx $(x) = - sin x - 2 (- l) , n2 n u 2 - 1

so that

f "(x) = - sin x - f(x)

f"(x) + f(x) = - sin r

Solving this differential equation for f(x) gives

f (x) = c1 cos X. + c2 sin x + +x cos r (9.10)

To find el, we set x = 0, obtaining f(0) = el. According to (9.8), f(0) = 0 so that cl = 0. To find c2, we differentiate (9.10) and compare the result with (9.9):

1 QD cos x x sin x cos IZX c2c0sx+- - 2 2 = - Z For x = 0, this gives

Thus we have sin x xcos x

f(x) = , + 2

. We now note another useful criterion for the differentiability of a function

defined by a trigonometric series:

THEOREM 3. Consider the series 00

+ 2 (- I)" (a, cos nx + bn sin nx), ' n = l

where a, and b, me positive. the quantities n a , nb. do not increase (Hter a certain n) and converge to zero as n -+ ao, then the series (9.1 1 ) converges for -n < x < n a d has a diflerentiable sunt f (x) for which

a0

f (x) = 2 (- lr n(b, cos nx - a, sin nx), (9.12)

i.e., the series (9.1 1 ) can be diflerentiated term by term.

Page 151: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Sac. 10 OPERATIONS ON FOURIER SERIES 1 37

Proof. By hypothesis, a,, and bn decrease monotonically to zero as n + oo. Therefore, by Theorem 1 of Ch. 4, Sec. 4, the series (9.1 1) and the series in the right-hand side of (9.12) converge uniformly on every interval [a, b] inside the interval [-x, x]. It follows that the series (9.11) can be differentiated term by term for -n < x < rr (see Theorem 2 of Ch. 1, Sec. 4), which proves (9.12).

Example 2. It is an immediate consequence of Theorem 3 that the series

has a differentiable sum for -7e < x < n, and that Q)

f ( x ) = - 2 (-1)" sin nx n u 2 In n

* 10. Differentiation of Fourier Series. The Case of a Function Defined on the Interval [0, n ]

As a simple consequence of Theorem 1 of Sec. 8, we obtain

THEOREM 1. Let f(x) be continuous and have an absolutely integrable derhative on [0, rr] (the derivative may not exist at certain points), and let f(x) be expanded in Fourier cosine series or Fourier sine series. Then the cosine series can always be differentiated term by term, while the sine series can be diarentiated term by term i f f (0) = f(n) = 0.

Proof. In both cases, the extension of f(x) onto [-z, 0] (which is even for the cosine series and odd for the sine series) leads to a function which is continuous on [-n, rr] and which takes equal values at the end points of [-n, x]. Therefore, in both cases, the subsequent periodic extension of f(x) onto the whole x-axis leads to a continuous function of period 2x with an absolutely integrable derivative. We can now apply Theorem 1 of Sec. 8.

THEOREM 2. Let f(x) be continuous and have an absolutely integrable derivative on [0, z] (the derivative may not exist at certain points), and let f(x) be expanded in the Fourier sine series

Then

Page 152: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

38 OPERATIONS ON FOURIBR SERIES

where

Proof. If QD ' + 2 a; cos nx, f (4 - 3

n = l

then

CHAP.

(10.2)

We know that

sin nx x n = l T

~ ~ i n & 2 + + ~ l ) x x 1 sin nx = s = z 2 I1 - (-1n- k=O n = l n (1 0.4)

(0 < x < X).

[See (13.9) and (13.11) of Ch. 1, Sec. 13.1 Therefore

(f'(x) - $) dx = f (x) - * - f (0) 0 2

QD PO

= 2 bnsinnx - 06 2 ( - l r+ ' sin nx

n = l n o 1 n 00 sin nx - Zf(0) 2 I1 - (- 7 n = l

a0 2 sin nx = 2 [nbn - If@) + (4 + ; f (~ ) ) ( - l~ ] 7

n u 1 X

which is the Fourier series of the function

But this series can also be obtained by integrating the series (10.3) term by term (see Sec. 7). and hence. conversely, (10.3) can be obtained by differentiating (1 0.5) term by term. Therefore

PO

f (x) - - 2 [nbn - 2f(0) + (ah + Zf(0))(-l)"] cos nx, n u 1 X X

Page 153: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 10 OPERATIONS ON F O ~ SERIES 1 39

which implies (10.1) and (10.2) if we set

then, instead of (10.1), we obtain

i.e., the Fourier series off'(%) Is obtained simply by term by term dlfer- entiation of the Fotuier series of f(x).

Proof: This case corresponds to the condition

already considend in Theorem 1. In fact, for even n, we immediately obtain c = 0. But then for odd n, we have - 2d = 0 or d = 0, and f(0) = f(n> follows from (10.2). Remark. Instead of (10.2), we can determine the constants c and

d by using the formulas

c = - lim nbn, where n is even, n+ ao

d = f( lim nb,, - c), where n is odd. (10.6)

To see this, we note that the Fourier cmfficients of the absolutely in- tegrable functionf'(x) converge to zero as n + ao. Therefore, it follows from (10.1) that

lim (nbn + C) = 0 woo

for even n, which implies the first formula (10.6), while for odd n

lim (nb, - c - 2d) = 0, woo

which implies the second formula (10.6). Theorems 1 and 2 have the following converses:

THEOREM 3. Suppose that we are given the series

Page 154: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

140 OPERATIONS ON FOURIER SERIES CHAP. 5

y the series10

- 2 nun sin nx

is the Fmier series of an absolutely integrable function cp(x), then (10.7) b the Fourier series of the fwction

which is continuow on [0, ~ 1 . 1 1 Moreover, (10.7) converges to this fwction, and obviously f (x) = p(x) at all the continuity points of ~ ( x ) .

This theorem is a simple consequence of Theorem 2 of Sec. 9, obtained by setting bn = 0 (n = 1,2, . . .).

TRPoR~M 4. Suppose that we are given the series Q 2 bn sin M. (10.8)

r=l

r f the limits (10.6) exfst and the series10

Is the Fourier series of an absolutely integrable fiction ~ ( x ) , then (10.8) S the Fourier series of the function

f (x) = p v(x) dk + (0 < x < z). 0

Moreover, (10.8) converges to thh fiction, and obvious& f'(x) = q(x) at all the continuity points of ~ ( x ) .

Proof. We can apply Theorem 2 of Sac. 7 to the series

obtaining

00 00 sin nx = 2 bnsinnx - 2 [l - ( - l r ]d

n=l n=1 ?a a0 7cd

= 2 bn sin nx - - n=l 2

10 I t is not assumed that this series converges.' 11 Since the sum of the series (10.7) is an even function, it will also be continuous on

[-sr, and hence on the whole x-axis.

Page 155: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

for 0 < x < x [see (10.4)]. Therefore

for 0 < x < x. As a special case of this theorem, we have the following result:

THEOREM 5. Given the series (10.8), if the limit

lim nb, = h n+co

exists, and i f the series

is the Fourier series of an absohitely integrable function ~(x) , then (10.8) Is the Fourier series of the f ic t ion

7th f(x) = lax 9(x) dx + p (0 < x < x).

Moreover, (10.8) converges to this function, and obviousZy f(x) = p(x) at all the continuity points of p(x).

To prove Theorem 5, we note that when the limit (10.10) exists, the formulas (10.6) give c = - h, d = h, and then the series (10.9) becomes (10.1 1).

Example 1. The series

is the Fourier series of a function which can be differentiated any number of times. In fact, in this case (10.10) gives

h = lim n4 = 1,

n4 + 1

and the series (10.1 1) is

1 - z + n=1 P(n:: - 1 ) cos nx

Page 156: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

This series converges absolutely and uniformly, and therefore has a con- tinuous sum ?(XI. By Theorem 5

n3 sin nx n = l

f'(4 = ?(XI

for 0 < x < n. Next, we note that the Fourier coefficients of the function p(x) satisfy

the inequality

Therefore, by Theorem 3 of Sec. 8, the function cp(x) has two continuous derivatives, and in fact

00

cpf(x) = 2 n sin nx n=1 n4 + 1'

We can now apply Theorem 3 to the last series, obtaining aB

n3 sin nx = - f(x). n=1

Obviously v*(x) = f (in)(x),

so that f(x) satisfies the differential equation

which implies that the derivatives of f(x) exist up to any order. As in Sec. 9, the results of this section can be used to evaluate the sums of

trigonometric series :

Example 2. Find the sum of the series

This series is uniformly convergent, and therefore has a continuous sum F(x). Differentiating the series term by term, we obtain

n sin nx (1 0.1 3)

Page 157: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 10 OPERATIONS ON FOURIER SERIES 143

We now apply Theorem 5 to this last series. Then, formula (10.10) gives

?82 h = li* (- ) = -1, -00 n2 + 1

and for the series (10.1 l), we obtain

l + f ( - n 2 + 1 + 1 cos nx 2 * = I "' )

Thus, according to Theorem 5, the sum f(x) of the series (10.13) is

X X f(x) = 2 + r F(x) dx - 3 (0 < x < x).

0

But then Theorem 3 is applicable to the series (10.12)' and therefore

1 F"(x) - F(x) = T

The solution of this differential equation is

F(x) = clex + c2e- - 1 2'

and hence

Ft(x) = c l r - c2e--x.

Setting x = 0 in-these expressions and using (10.12) and (10.14), we find that

from which we obtain

Then, with these values of cl and c2, the function (10.15) gives the sum of the series (10.12).

Page 158: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

1 44 OPEUTIONS ON FOURIER SERIES CHAP. 5

I I. Improving the Convergence of Fourier Series

In the applications, the most convenient trigonometric series are those with rapidly decreasing coefficients. In this case, the first few terms of a series sufEce to give its sum quite accurately, since the sum of all the remaining terms will be small if the coefficients approach zero rapidly enough. Thus, the faster the coefficients decrease, the fewer terms of the series are needed to give its sum to any degree of accuracy. It should also be noted that in differentiating trigonometric series, the situation is the simplest when the series have rapidly decreasing coefficients (see Theorem 3 of Sec. 8). These considerations lead us quite naturally to the following problem:

Suppose we are given a trigonometric series

a0

f(x) = 2 + 2 (an cos nx + bn sin nx), nu1

whose sum we denote by f(x). How can we,subtract from this series another trigonometric series whose sum ~ ( x ) is known (in finite terms) in such a way that the series which is left, i.e., the series related to f(x) and cp(x) by the formula

a0

f(x) - p(x) = 2 (a,, cos nx + p, sin nx), n= 1

has quite rapidly decreasing coefficients? Once this problem has been solved, then operations on f(x) can be replaced by operations on the known function ~ ( x ) and on a series with rapidly decreasing coefficients.

The fact that this problem can be solved in cases of practical interest can be made plausible by the following argument: Suppose that we are given a function f (x), defined on [-x, x ] (or on [0, xJ), and suppose that f (x) has several derivatives. Then, the periodic extension of f(x) over the whole x-axis may lead to a discontinuous function (or to a function with discon- tinuous derivatives), with the result that the Fourier coefficients fall off slowly. It is not hard to see that by subtracting a suitably chosen linear function from f(x), we can convert f(x) into a function with equal values at the end points of the interval [-x, n], and hence into a function which can be extended continuously over the who1e.j-axis, i.e., into a function with Fourier coefficients which fall off faster than those of the original function. Similarly, by subtracting a suitably chosen polynomial from f(x), we can arrange for not only the function but also several of its derivatives to have equal values at the end points of the intervd. But then both the function and these derivatives can be extended continuowly over the whole x-axis, and Theorem 2 of Sec. 8 can be applied. This already guarantees that the coefficients fall off quite rapidly.

Page 159: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 11 OPERATIONS ON FOURIER SERIES 145

Thus, there is hope for solving our problem. However, in the actual problem, we are given a series and not a function. Therefore, the form of the function ~ ( x ) has to be inferred from its series expansion; this constitutes a difficulty which, however, can very often be overcome. When the problem has finally been solved, we say that we have improved the convergence of the series ( 1 1.1).

We now show two ways of improving the convergence of Fourier series. The first method is based on the fact that the difference between two infinites- imals of the same order is an infinitesimal of a higher order than the order of either of the original infinitesimals.12

Example 1. Improve the convergence of the series 00 n3

fix) = 2 ( - l Y n 4 - I sin nx. n=2

In this case, the quantity n3/(n4 - 1 ) is of order l/n as n + oo [since n3/(n* - 1 ) + lln = n4/(n4 - 1 ) + 1 as n -* oo]. A simple calculation shows that

Therefore we have 00 sin nx 00

f(4 = 2 (-1)" sin nx

+ C ( - 1 ~ ~ s - n o 2 n n=2

But according to formula (13.9) of Ch. 1 , Sec. 13

sin nx x n 2 (-x < X < X),

n-1

and hence

X Q)

+ sin x + 2 (- 1)" sin nx f(x) = - - 2 ( - x < X < a.

n = 2 ns - n

In the last series, obviously

Ib,#I < M (M = const),

i.e., the Fourier coefficients are of order l/n?

Example 2. Improve the convergence of the series

f (x) = 5 n4 - n 2 + 1

nZ(n4 + 1 ) cos nx. n = l

-

12 Two infinitesimals are said to be of the same order if their ratio converges to 1.

Page 160: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

146 OPBMTIONS ON PO- SBRIES CHAP. 5

In this case, the Fourier coefficients are of order l/n2 and

Therefore we have

COS m QO

f(x) = n2 cos nx

n= 1 - n = 1 C n 4 + 1.

But according to formula (13.8) of Ch. 1, Sec. 13

cos nx 3x2 - 61cx + 2x2 = 12 (0 < x < 2%),

n= 1

and hence

In the last series

1as41 G 1, i.e., the Fourier coefficients are of order 1/12?

The second method of improving convergence13 is based on representing the Fourier coefficients as sums of the form

(A = const, B = const,. . .). Example 3. Improve the convergence of the series

PO sin nx f ( ~ ) = 2 + a (a = const, a > 0).

n = l

Obviously we have14

We truncate the series in parentheses at the term a2/n2, and then add up the remainder, obtaining

1 a3 1 a a* = - - - + - - a3 n + a n n n2 nqn + a)) n n2 n3 n3(n + a).

13 Actually, it can be Shown that the second method reduces to repeated application of the first method.

14 Of course, we can only talk about an infinite series if a/n < 1, but the result we need is always true, as can easily be verified.

Page 161: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Therefore 00 sin nx a0 sin nx QO sin nx PO sin nx

f ( ~ ) = 2 -7 - a C 7 + 2 n3 - a3 2 n3(n + a)s n = l n = l n u 1 n=1

and the sums of the first three series, and more generally the sums of series of the form

(where p is a positive integer), can easily be obtained by using known ex- pansions. In fact, according to formulas (13.7), (13.8) of Ch. 1, Sec. 13, and (14.1) of Ch. 3, Sec. 14

s cos nx 3x2 - 67cx + 2trz

2 = - ln (2 sing) (0 < x < 2 ~ ) . n= 1

Integrating the second and the third series, we obtain

2 siY = - I~~ ~n (2 sin ;) dx n u 1

for 0 < x < 25~. Therefore

a2 QO sin nx + (n3 + 3 ~ x 2 - 2x2~) - a3 2 n3(n - ay n-1

where the coefficients of the last series are of order l/n?

12. A List of Trigonometric Expansions

In dealing with Fourier series, it is convenient to have a list of commonly encountered series. Such a list is particularly useful if we are trying to improve the convergence of a series. In the list given below, we have gathered together all the expansions obtained previously, and we have added some new ones.

Page 162: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

148 OPERATIONS ON POWER smms CHAP. 5

COS I1X 1) 2-= n - 1n (2 sin ;) (0 < x < 279, [Ch. 3, (14.1) ] n = 1

2) 2 s i y - n - x - 2 (0 < x < 279, [Ch. 1, (13.7)]

n u 1

cos nx 3x2 - h x + 2x2 12 (0 g x g 2x), [Ch. 1, (13.811

n= 1

= - [ 1n (2 sin ;) dx (0 g x < 2x), [See Sec. 1 I] n= l

5) 5 m:y = IOx ti* jox ln (2 sin ;) d~ + 2 a - n3 1 (0 g x g 274 n o 1 n = l

n3 (Zl ; = 25.79436. . . = 1.20205.. .), which is obtained by term by term integration of the preceding series,

2 si;nx - x3 - 37& + 2 x 2 ~ - 12 (0 g x g k), [See Sec. 111

n = 1

7) 2 (-lY+l cos nx n = 1. (2 ms i) (-n c x < n), [Ch. 3, (14.2)]

n = 1

8) 2 (-ly+l sin nx x = -

2 ( - x < X < X), n [Ch. 1, (13.9)J n= 1

10) 2 (-ly+l sin nx = 1 in (2 cos i) dx n2

(-x $ x G n), n= 1

obtained by term by term integration of the series 7).

11) 5 (-ly+l cos nx n= 1 n3

00 1 = ~ ( - l ~ + 1 - - $ d x ~ l n ( 2 c o s Z dr ( - x g x g n ) ,

n=1 n3 3 obtained by term by term integration of the series lo),

sin nx 7e2x - x3 12) 5 (- 1y+l- - n3 - 12 {-n G x < x),

n u 1

Page 163: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

obtained by term by term integration ofthe series 9).

00s (2n + 1)x 1 x = - zlntan2 (0 < x < st),

n=O

obtained by addition of the series 1) and 7),

sin (2n + 1)x st = z (0 < x < st),

n=0 [Ch. 1, (13.11)]

cos (2n + 1)x n2 - 2nx - - 8 (0 g x < n), [Ch. 1, (13.12)J

n=O

sin (2n + 1)x 1 x x = - t i l n t a n Z d ~ ( O g x g x ) ,

n=O

obtained by term by term integration of the series 13),

cos (2n + 1)x 17)

n=0 f (2n + 1)3

1 Po

= Zkxdxrlntan;dx o + 2 (2n + 1 113 (0 < x < x), n=O

obtained by term by term integration of the series 16),

sin (2n + 1)x x2x - zx2 18) 5 (2n + 1)s

0 - 8 (0 < x ,< x),

n = 0

obtained by term by term integration of the series 15). If in the formulas 13) through 18), we replace x by t and then set t =

+z - x, we obtain the expansions

19) 5 (-1)" cos (2n + 1)x - x 2 n + 1 - a ( < x < " ) 2 2 ' n=O

20) f ( - l y sin (2n + 1)x 1 2n + 1 ' - 2 ln tan (: - $)

n=O

21) 3 (-lyJ cos (2n + 1)x t (2n + 1)2

= - $ r-x ln tan dt n o 0

22) S ( - l y sin (2n + 1)x xx =- (2n + 1)2 4 n=O

Page 164: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

I SO OPBMTIONS ON FOURIER SERIES CHAP. 5

24) $(-lr sin (2n + 1)x dr r in tan d ! (zn + 113 n o 0 0

13. Approximate Calculation of Fourier Coefficients

In practical problems, functions which have to be expanded in Fourier series are often given as tables or graphs, i.e., approximcrtely, rather than analytically. In this case, the Fourier series cannot be obtained by direct application of the usual formulas

and we are faced with the problem of calculating a,, and bn approximately. (In most cases, it is sufficient for practical purposes to know only tne first few Fourier coefficients.) To solve this problem, we go from the exact formulas (13.1) to approximate formulas, by using the technique of approx- imate integration. The usual method is to use the rectangular rule or the trapezoidal rule. In our case, the rectangular rule reduces to the following procedure:

Let the interval [0,2nJ be divided into m equal parts by the points

and suppose that the values of the function f (x ) at these points are known to be

YO, Y1, Y29 - 9 Ym-1, Ym*

Then we have

2 m-1 2 h b, z - 2 y, sin - m n, k=O

where x denotes approximate equality. For example, suppose m = 12.

Page 165: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

a. 13 OPERATIONS ON FOURIER SERIES I 5 I

Then, the numbers (13.2) are

or in degrees

In this case, it is easy to see that all the factors multiplying the ordinates in (13.3) reduce to

0, f 1, f sin 30" = f 0.5, f sin 60" = f 0.866,

and it is easily verified that

6 a ~ YO + Y1 + 7 2 + Y3 + Y4 + YS + y6 + Y7 + y8 + y9 + Yl0 + Yl1, Ql b 0 - Y S ) + o*866bl + Yl l - Y S - ~ 7 ) + 0.5@2 + Yl0 - Y4 - ~ 8 ) s

6a2 6'0 + Y6 - Y3 - ~ 9 ) + o**i + ys + 77 + Yl l - y2 - Y4 - Y8 - ~ 1 0 ) ~ e3 7 0 + Y4 + Y8 - 7 2 - Y6 - Yl0,

o*Ql + 7 5 - 7 7 - 7 1 1 ) + 0.86602 + y4 - y8 - ~ 1 0 ) + 6 ' 3 - ~ 9 ) .

6b2 a 0*866@l + y2 + y7 + y8 - y4 - y5 - 710 - yll) , 9 3 = Yl + Ys + Y9 - Y3 -- Y7 - Y11. etc. (1 3.4)

To simplify the calculations, it is convenient to perform them according to the following scheme: First we write the ordinates yo, yl, y2, . . . in the order shown below. Then we form the sum and difference of each pair of ordinates such that one ordinate appears below another. (The number zero is understood to appear in positions where there is no entry.)

Then we write down these sums and differences and form new sums and differences in a similar way:

Sum Difference

YO Y1 Y2 Y3 Y4 YS Y6 Y11 Yl0 Y9 Y8 Y7

uo ul u2 u3 u4 us u6 0 1 02 V3 04 vs

Sum Difference

uo U l u2 u3

u6 U S u4

so 81 82 83

to tl t 2

Sum Difference

v1 v2 v3 as 04

01 0 2 0 3

T I .rz

Page 166: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

CHAP. 5

Using these quantities, we can write

620 = so + Sl + 82 + sj, 6al c to + 0.866tl + 0.5t2, 6a2 z SO - 83 + 0.5(s1 - s2), 6a3 NN to - 12,

6bl z 0 . 5 ~ ~ ~ + 0.866~~ + 0 3 ,

6b2 z O.866(q + 4, 6b3 z a1 - 03,

instead of (13.4). We have given a scheme for carrying out the calculations when twelve

ordinates are given. If this scheme is used for smooth functions for which we know the exact values of the Fourier coefficients, we find that the resulting approximate values of the coefficients ao, al, bl, a2, q, ah b3, are quite close to the exact values. To obtain more exact results, and also in cases where a larger number of Fourier coefficients are needed, we can use a scheme with a larger number of ordinates. The scheme with 24 ordinates is frequently used.

PROBLEMS

1. Calculate the sums' of the following series :

Hint. Use the list in Sec. 12 and Parseval's theorem (Sec. 3).

2. Calculate the following integrals :

a) 1 ln2 (2 sin i) h,

Hint. Use the list in Sec. 12 and Parseval's theorem. 3. Suppose that the function f(x) is continuous and has three continuous deriva- tives on the interval [0, lc], and f(0) = f(=) = 0, fa(0) = f "(=) = 0. Show that the Fourier series of f(x) can be differentiated twice term by term, and show that the series

converges,

Page 167: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

PROBLEM OPERATIONS ON FOURIER SERIES 1 53

4. A function is defined by the series

c08m sin nx ( x ) = 2 (- +

n o 1 n3 n + 1

Write the Fourier series of its derivative. Hint. Use Theorem 2 of Sec. 9.

5. Show that the function defined by the series QO sin nx

n= 1

is differentiable, and write the Fourier series of its derivative.

Hint. Use Theorem 3 of Sec. 9.

6. Show that each of the following functions is differentiable, and write the Fourier series of its derivative:

n + l a) AX) = 2 n2 + n + 1 sinnx ( ~ < x < x ) ;

n=1 00 n2 sin ax (0 < x < =).

Hint. Use Theorem 5 of Sec. 10.

7. Find the sum of the series aD

2 ,S"-"Iw n=2

Hint. Proceed as in Example 2 of Sec. 10.

8. Improve the convergence of the following series:

n 3 + n + 1 OD n3

n(n3 + 1) sinnx; b) z n , + l s i n a x ; n= 1 n = l QD

C) 2:: (In this example and those that follow, a > 0); n = l

a0 sin nx 1 ~(-l)n-t-lW2-

n = l n + a' n=1 n + a'

sin (2n + l)x. 3 n + a , n u 1

Hint. In a) and b), proceed as in Examples 1 and 2 of Sec. 11. In c), d), and e), use

a2 n nqn + a)'

Page 168: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

154 OPBMTION~ ON wxrrug~ SERIES CHAP. 5

divide each series into three series, and use the list in Sec. 12. In f ) and g), use

divide each series into three series and use the list in Sec. 12.

9. Show that if f(x) is square integrable and if , QD

then 1 x+h QO sin nh

/x-h f(u) ckc = 3 + 2 (a,, cos nx + bn sin a) 7

n= I

where Ih( 9 x.

10. Show that if f(x) has a square integrable derivative f'(x) and if

then

11. Let f(x) have period 2% and Fourier coefficients a, and 6,. Show that

where p i = 4 + bj,. Furthermore, show that iff (x) satisfies a Holder condition of order a, i.e., if there are positive numbers c and a such that

If(4 - nY)l C clx - yla for all x and y, then

where co > 0. (0. Ch. 1, Rob. 7.)

12. Let f(x) sat@ all the conditions of the preceding problem. Show that

Use c) to prove that if a > 3, then the Fourier series of f(x) converges absolutely Ukrnstein's theorem).

Page 169: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SUMMATION OF TRIGONOMETRIC FOURIER SERIES

I. Statement of the Problem

Given a trigonometric Fourier series

Q,

+ 2 (a,, cos nx + bn sin nr) 2 n = i

about -which we know only that it is the Fourier series of some function f(x), can we find the function f(x)? If we know in advance that the series (1.1) converges to f(x), then we can obtain f(x) as the limit of the partial sums of the series. The situation is different when we have not succeeded in proving that the series converges or when the series turns out to be divergent, for then we either do not know whether or not the partial sums have a limit or else we actually know that the limit does not exist. Thus, we have to find an operation which allows us to detennine a function from a knowledge of its Fourier series, regardless of whether or not the series converges. This is the problem that will concern us in this chapter. The operation in question will be referred to as summation of the series. Summation must not be con- fused with the operation of finding the sum of a series which is known before- hand to converge, since summation can also be applied to divergent series.

The problem of summation can also be posed for arbitrary series of numbers or functions. Naturally, a properly defined summation operation

1 55

Page 170: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

must always have the property that it gives the sum of the series in the usual sense, if the series converges.

2. The Method of Arithmetic Means

Consider the series

u0 + Ul + U2 + * * * + un + * * * ,

and let

It may happen that whereas the series (2.1) diverges, the quantities o, (the arithmetic means of the partial sums of the series) converge to a definite limit as n + m. For example, the series

diverges, but in this case so = 1, SI = 0, sz = 1, sj = 0,. . ., SO that an = f. for even n and an = % + (2n)-1 for odd n, and hence

lim a, = +. n-00

More generally, if

lim a, = a, n+m

we say that the series (2.1) is summable by the method of arithmetic means1 to the value a.

We now ask whether this method of summation satisfies the requirement mentioned at the end of Sec. 1, i.e., does the number a give the sum of the series (2.1) in the usual sense, if the series converges? The answer is in the affirmative, as the following theorem shows:

THEOREM. If the series (2.1) converges and its sum equals 4 then the series is summable by the method of arithmetic means to the same d e r o.

Proof. Sincc by hypothesis

lim s, = a, I)--+=

then for any r > 0, there exists a number m such that

1 Also known as Cesdro's method of summation. (Translator)

Page 171: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 2 SUMMATION OF TRIGONOMfZTRIC FOURIER SERIES 1 57

provided that n 2 m. Now consider

Forn > m

1 m - l 1 n-1

a n - 0 ' -

and hence

Since m is fixed

for all sufficiently large n. On the other hand, by (2.2) 1 n - I n - m - l c E - 2 Isk - 01 <

,=, n 2 < 2' Therefore

1% - 01 < c

for all sufficiently large n, which proves the theorem.

Of course, the method of arithmetic means is also applicable to series of functions, in particular, to Fourier series.

3. The Integral Formula for the Arithmetic Mean of the Partial Sums of a Fourier Series

Suppose that PO

f ( x ) - + (ak cos kx + bk sin kx), k = 1

Then for the arithmetic mean of the partial sums, i.e., for

s0(x) + sl(x) + + sn- (x) %(XI = 9 n

Page 172: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

we obtain

a 0 n- 1 n - k

a&)=?+ 2 y (ak cos kx + bk sin kx). k=1

We can also write an integral formula for on(x). In fact, as we know from (4.1) of Ch. 3

sin (n + +)u

aria hence

= 1 1" f(7 + sin (L + +)u du. * 2 810. (~12)

We now calculate the sum of sines appearing in this expression. Since

U 2 sin 2 sin (k + +)u = cos ku - cos (k + l)u,

we have n- l n-1

2 sin; 2 sin (k + 4)u = 2 (cos ku - cos (k + 1)u) k - 0 k = 0

nu = 1 - cos nu = 2 sin2 - 2 '

so that

Therefore

which is the desired integral formula. We note the following consequence of (3.3). Suppose that f(x) = 1 for

al lr Thensn(x) = 1 (n = 0, 1.2 ,... ),andhenceo,(x) = 1 (n = 1,2 ,...) so that (3.3) implies

4. Summation of Fourier Series by the Method of Arithmetic Means

THEOREM 1. The Fourier series of an absolutely integrable function f(x) of period 2z is sumrnable by the method of arithmetic means to f(x)

Page 173: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 4 SUMMATION OF TRIGONOMETRIC FOURIER SBRIES 1 59

at every p o w of continuity and to the value

at every point of jtunp discont&ufty.

ProoJ Since

at continuity points, it is sufficient to prove the relation

and to show this, it is in turn sufficient to prove that

[sa (3.3)]. Both of these formulas are proved in the same way, and hence we shall consider only the first of them. Since the integrand of (3.4) is an even function, we have

so that

Thus, it follows from (4.1) that we have to prove the formula

Let r > 0 be arbitrary. Then, since

lim f(x + u) = f(x + O), n-bx U> X

it follows that

Page 174: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

1 60 SUMMATION OF TRIQONOMETRIC FOURIER SERIES CHAP. 6

for 0 < u ( 6, if 6 > 0 is sufficiently small. We now divide the integral appearing in (4.4) into two integrals:

Then (4.5) implies that

whence, by (4.3) G

11'1 < 2 for any n. On the other hand

and therefore, for all sufficiently large n

G 1121 < T (4.8)

The formula (4.4) follows from (4.6), (4.7) and (4.8).

THEOREM 2. The Fourier series of an absolutely integrable fMclion f (x) of period 2n is uniformly summable by the method of arithmetic means to f (x) on every interval [a, 6 1 lying entirely within an interval of con- tinuity [a, b] o f f (x). Uniform summability on [a, p] means that for any E > 0 and any x in

[a, 61, there exists a number N such that

If ( 4 - o.(x)l < c provided that n 3 N.

Proof. Let x lie in the interval [a, $1. Then, using (3.3) and (3.4), we can write

where J denotes the integral from 0 to n, and j denotes the integral from -IT to 0. Let 8 > 0 be so small that

Page 175: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

for any x in [a, PI, provided that lul < 6. [It was to guarantee the existence of such a 6 that we required that the interval [a, P ] should lie entirely within a larger interval [a, b] of continuity of f(x).]

Now let M denote the maximum of I f(x)l on [a, PI, and consider

Then, by (4.10), the inequality

holds for any x in [a, PI. [See the proof of the inequality (4.7).] On the other hand, for any x in [a, PI, we have

Here, the term in parentheses is a constant, and hence there exists an N such that the inequality

holds for all n 2 N. But then, by (4.1 1)

for all n 3 N and any x in [a, 61. A similar inequality can be proved for the integral j, by using the

same method. The proof of the theorem is then completed by using (4.9).

Theorem 2 has the following remarkable consequence:

THEOREM 3. The Fourier series of a continuous function f (x) of period 2x is unfformly summable to f(x) by the method of arithmetic means.

To emphasize the power of the theorems just proved, we recall once more that the Fourier series of even a continuous function f(x) may be divergent, so that the partial sums of the Fourier series may be bad approximations to Ax) . However, if f(x) is continuous, the arithmetic means of the partial

Page 176: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

sumss i.e., the sums of the form (3.1) will be unflorm approximations to f (XI*

The theorems just proved can be used to make certain questions con- cerning the convergence of Fourier series more precise. Thus, Theorem 1 and the result of Sec. 2 imply

THW)REM 4. the Foralcr series of an absolutely integrable k c t i o n f (x) converges at a point of continuity o f f (x) (or at a point of jump dis- continuity), then its sum must equalf(x) (or +W(x + 0) + f(x - 0)D.

THW)REM 5. 11 the Fourier series of an absolutely htegrablefMction f(x) converges everywhere, excebt possibly at a finlte number of potrts, then its sum equakk f(x), except possibly at certain points.

Theorem 5 reduces to Theorem 4 if we recall that an integrable function, as we understand the tenn, can be discontinuous only at a finite number of points.

Finally, Theorem 1 also implies that given the Fourier series of an absolutely integrable function f(x), the method of arithmetic means can be used to reconstruct f(x) at all its continuity points, i.e., everywhere except possibly at a finite number of points. Therefore, we have the following important proposition, which generalizes Theorem 3 of Ch. 5, Sec. 3:

THEOREM 6. Any absolutely integrable fMction ih completely defined (except for its values at a finite number of points) by its trigonometric Fourier series, whether or not the series converges.

5. Abel's Method of Summation*

Consider the series

and also the series

We assume that the series (5.2) converges for 0 < r < 1 [which will always be the case if the terms of the series (5.1) are bounded] and that the limit

lim d r ) = a r+1

exists, where o(r) is the sum of (5.2). In this case, we say that the wries (5.1) is sununable by Abel's method to the value a.

2 Equivalently, the "method of convergence factors" - Russian "M~TOCI ~~ MH-a." (lhlslaror)

Page 177: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Abel's method can be used to sum certain divergent series. For example, the series

already encountered in Sec. 2 is summable by Abel's method to the value o = .), as well as by the method of arithmetic means. In fact, in this case

and therefore lim a(r) = +. -1

We now ask whether Abel's method of summation gives the sum of the series (5.1) in the usual sense, if the series converges. The answer is in the affirmative, as the following theorem shows:

THEOREM. If the series (5.1) converges and its sum equals 0, then the series b summable by Abel's method to the same number a.

Proof. If the series (5.1) converges, then by the Lemma of Ch. 4, Sec. 6, the series (5.2) converges and its sum a(r) is continuous on the interval 0 s r d 1. This means that

which proves the theorem.

6. Poisson's Kernel

We now calculate the sum of the series

To do this, we consider the series

+ 2 fl, = r(cos cp + i sin cp). Z n u l

Since Izl = r < 1

1 00 1 z + 2 z n = Z + - = l + z - 1 + rcosp + irsincp Z n= I 1 - z 2(1 - z) - 2(1 - r cos cp - ir sin cp)

- - (1 + r cos p + ir sin p)(1 - r cos p + ir sin cp) 2[(1 - r cos cp)2 + r2 sin2 cp]

- - 1 - r2 + 2ir sin cp 2(1 - 2r cos cp + rz)'

Page 178: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

On the other hand

Therefore

1 00 1 + 2 FCOsnq = - 1 - r2

3 n= 1 2 1 -2rcoscp+r2 (0 d r < 1), (6.1)

and at the same time we have also obtained the formula PO r sin p 2 F sin ncp =

1 1- Zcoscp + r2 (0 g r < I). n = l

The function

of the variables r and p is called Poisson's kernel. It should be pointed out that Poisson's kernel is a positive quantity, since

7. Application of Abel's Method to the Summation of Fourier Series

Let f(x) be an absolutely integrable function, and let 00

f ( ~ ) - + 2 (an cos nx + b, sin n ~ ) . nu1

Consider the series 00

.f(& r) = 3 + 2 F(O. COS n* + bn sin PIX), n= l

where 0 < r < 1. This series converges, since an + 0 and bn + 0 as n + 00, and therefore

141 GM, lbnl <A4 ( n = 1 , 2 ,...; M - c o ~ s ~ ) ,

IF(u,, cos nx + bn sin nx)l ( ~ M P ,

where 2Mr*r is the general term of a convergent series, since r < 1. If

Page 179: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 7 SUMMATION OF TRIOONOMETRIC FOURIER SERIES 1 65

Lim f(x, r) exists, this means that the series (7.2) is summable by Abel's r+l

method. For convenience in studying the properties of this kind of summation,

we first represent the function f(x, r) as an integral. Recalling that

we can write

But for fixed r < 1 and x, the series

converges uniformly in t (since its terms do not exceed in absolute value the corresponding terms of the series

which is known to converge) and can therefore be integrated term by term. Thus, the series

QO

fo + 2 ~ f ( t ) cos n(t - x) 2 n = l

can also be integrated term by term, and instead of (7.3), we can write

if we use (6.1). In this way, we have written f(x, r) as an integral known as Poisson's integral. It should be noted that if f(x) = 1, then ao/2 = 1, an = 0, bn = 0, and hence f (x, r) r 1, SO that (7.4) becomes

Page 180: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

THEOREM 1 . Let f (x) be an absolutely fntegrable function of period 2n. men

at every point where fix) is continuous, and

lim f(x, r) = f(x + 0) +f(x - 0) ?-+I 2

at every point where f(x) has ajump discontinuity. In other words, the Fourier series of Ax) is summable by Abel's

method to the value f(x) at every point of continuity of f(x) and to the value +lf(x + 0) + f (x - O)] at every point of jump discontinuity of f(x).

Pr04f. S t t - x = u in (7.4). Then

or, since the integrand is periodic

1 - r2 du.

Similarly, we can write

instead of (7.5). Since

at the continuity points of f(x), it is sufiicient to prove that

lim f (x, r) = f(x + 0) +f(x - 0) -1 2

at every point where the right-hand and left-hand limits exist, and to show this, it is in turn sufficient to prove the formulas

Both of these formulas a n proved in the same way, and hence we shall consider only the first of them.

Page 181: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Since the integrand in (7.7) is even (in u), the relation

holds, so that

f(x + 0) 1 = - f(x + 0)

1 - r2

2 2n o 1 - 2rcosu + r2 du.

Therefore, instead of (7.8), we can prove the formula

Let E > 0 be arbitrary. Then, since

it follows that

for 0 < u < 8, if 8 > 0 is sufiiciently small. We now write the integral in (7.10) as the sum of two integrals

Since Poisson's kernel is positive, it follows from (7.11) that

or, if we use (7.9)

for any r (0 < r < 1). On the other hand

[see (6.3)]. Noting that

lim 1 - r3 = 0

r+l r for all r near enough to 1, we find that

The formula (7.10) now follows from (7.1 2), (7.1 3) and (7.1 5).

Page 182: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

168 S ~ ~ ~ T I O N OF TRIGONOMETRIC FOURIER SERIES CHAP. 6

THEOREM 2. The Fourier series of an absolutely integrable function f(x) of period 27t is unifomly summable by AM'S method to f(x) on every interval [a, p] lying entitely within an interval of continuity [a, b] o f f (XI*

Uniform summability on [a, p] means that for any o > 0 and any x in [a, p], there exists a number ro (0 < ro < 1) such that the inequality ro < r < 1 implies the inequality

Proof. Because of (7.7), we can write

where J is the integral from 0 to x, and j is the integral from - x to 0. We can choose 8 > 0 to be so small that

for any x in [a, PI , provided that lul < 8. Now consider

By (7.18), we have

for any x in [a, 1 [cf. the proof of the inequality (7.13) 1. On the other hand, for any x in [a, p] we have

[d. (7.14)], where 1 f(x)l < M = const, for a < x ( [Recall that f(x) is continuous on [a, P I ! ] Since the term in parentheses is a constant, there exists a number ro (0 < ro < 1) such that the inequality ro < r < 1 implies that

Page 183: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 7 SUMMATION OF TRIQONOMETRIC F O U R I ~ SNES 1 69

for any x in [a, PI. But then

forro < r < 1,a < x < b. A similar inequality can be proved for the integral j, and then (7.16)

follows from (7.17), which proves the theorem.

Theorem 2 implies the following result:

THEOREM 3. If f(x) is a continuous function of period 2x, then f (x, r) + f (x) as r -, 1, uniformly f ~ r all x.

In other words, the Fourier series of a continuous function f(x) of period 2n is uniformly summable by Abel's method to f(x).

We now cite without proof the following remarkable theorem:

THEOREM 4. If an absolutely integrable fmction f (x) of period 2x has a derivative f (m)(x) of order m at the point x, then the series obtained by dzrerentiating the Fourier series of f(x) m times term by term is summable by Abel's method to the value f (m)(x).

It should be noted that when we differentiate a Fourier series term by term, we do not in general obtain the Fourier series of the derivative, even if the derivqtive exists for -7c < x < n. Moreover, as a rule, term by term differentiation gives series whose coefficients do not approach zero (and such series can be shown to diverge) or even approach infinity. Thus, for example, we know that

X Q)

= 2 (-1y+' sin nx (-7c < x < n).

1 , ,=l n

Term by term differentiation of this series gives

Po

(-l)"+' cos nx, (7.19) n= l

and another differentiation gives

- 2 (-ly+' nsinnx. n= 1

(7.20)

According to Theorem 4, the series (7.19) is surnmable to + for -x < x < x, and the series (7.20) is summable to 0.

Thus, while from the standpoint of ordinary convergence, the legitimacy of term by term differentiation'of a series can only be guaranteed by rather strong requirements, Theorem 4 shows that from the standpoint of Abel's

Page 184: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

method of summation, term by term differentiation of a Fourier series (a number of times equal to the differentiability of the function itself) is always legitimate, even if the resulting series turns out to be divergent.

PROBLEMS

1. Sum the following series by the method of arithmetic means:

Hint. For a) use formula (3.2) of Ch. 6, for b) use formula (2.1) of Ch. 4.

2. Show that Abel's method of summation gives the same answer as the method of arithmetic means for the series in Rob. 1.

3. Show that for 0 < r < 1, the two series

can be diffe~entiated term by team any number of times with respect to r and cp.

4. Sum the following series by Abel's method of summation:

a) 1 - 2 + 3 - 4 + - o o ,

b) p - 2pZ+ 3p3- 4 f l+ - - * , Ipl < 1 .

Hint.

S. Calculate the sum of the convergent Series

Why is this series convergent?

Hint.

Page 185: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

PROBLEMS SUMMATION OF TRI~ONOMETRIC FOURIER SERIES 17 1

6. Let f(x) be square integrable, with Fourier series

2 + 2 (a- kx + bk sin kx), k=1

and let an(%) be the arithmetic mean of the partial sums of the Fourier series of f(x), as in Sec. 3. Prove that

7. Sum each of the following series by the method of arithmetic means and also by Abel's method:

C) 1 - 2 + 3 - 4 + 5 - 6 + 7 - * . * (cf. hob. 4a) ;

8. With the notation of Sec. 2, show that if uo + ul + uz + + un + is summable by the method of arithmetic means, then sn/n + 0.

9. Let uo + u, + u2 + + Un+ be summable by themethodof arithmetic means, and let tn = ul + 2u2 + + nun. Show that

a) The series uo + ul + u2 + + Un + is convergent if and only if tn/n -+ 0;

b) If nu,, -+ 0, then the series uo + ul + u2 + + u, + is convergent.

10. With the notation of Sec. 2, show that

11. Let 5 un be summable to the value S by the method of arithmetic means, n=O

and let

n=o

Show that 00

I ~ ( x ) - S = (1 - x ) ~ 2 (n + l)(an+l - S)x(l. n=O

Page 186: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

I fZ SUMMATION OF TRIGONOMETRIC FOURIER SERfES CHAP. 6

Now let s > 0 be given, so that by assumption, there is an integer N such that if n 3 N, then laHr - Sl < c. Show that

and hence that

Comment. It follows that a series which is summable by the method of arithmetic means is summable to the same value by Abel's method.

Page 187: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

DOUBLE FOURIER SERIES. THE FOURIER INTEGRAL

I. Orthogonal Systems in Two Variables

Let R be a rectangle in the xy-plane described by the inequalities a ( x ( b, c < y g d, and let

~pn(x,y) ( n = 4 1 , Z * * * ) , (1.1) be a system of continuous1 functions defined on R, none of which vanishes identically. The system (I. 1) is said to be orthogonal if

provided that n # m. The number

is called the norm of the function qn(x, y). The system (1.1) is said to be normalized if

IIVnII-1 (n=O,1,2,***) or equivalently

1 Instead of continuous functions, we can also consider square integrable functions, as in Ch. 2.

173

Page 188: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

174 DO= PO- SERIES. THE FOURIER MTEORAL CHAP. 7

Every orthogonal system can be normalized, i.e., constants (n = 0, 1.2, . . .) can always be chosen such that the new system of functions

which is obviously still an orthogonal system, is also normalized. In fact, it is suflscient to set

Just as in the case of one variable (see Ch. I), we can associate a Fourier saies with every absolutely integrable function f(x, y) defined on R, i.e.,

where

In the case where the eqzuzlity holds in (1.3) and the series on the right con- verges uniformly, we find the expression (1.4) by multiplying (1.3) by each of the (continuous) functions 'pn(x, y) and integrating term by term. The quantities cn given by (1.4) are called the Fwier coeflcients of f(5 y).

In approximating any square integrable function f(x, y) by a linear combination of functions of the system (1.1), we find that the Fourier coefficients give the least mean square error, in the way described for fuctions of one variable in Ch. 2, Sec. 5. Moreover, we also have Bessel's inequality

where if the equality sign holds for any square integrable function, the system (1.1) is said to be complete. All the properties of complete systems proved for functions of one variable in Ch. 2, Secs. 7 and 8, are still valid. The completeness criterion of Ch. 2, Sec. 9 is also valid (properly rephrased for the two-dimensional case, of course).

The reader who has carefully read this section and Ch. 2 will s6e clearly how to generalize all that has been said about orthogonal systems to the case of functions of any number of variables.

Page 189: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 2 DOUBLE FOURIER SERIES. THE FOURIER INTEGRAL 175

2. The Basic Trigonometric System in Two Variables. Double Trigonometric Fourier Series

The functions

1, cos mx, sin mx, cos ny, sin ny, . . ., cos mx cos ny, sin mx cos ny, (2- 1) cos nuc sin ny, sin mx sin ny, . . . (m= 1,2 ,...; n = 1,2 ,...)

form the basic trigonometric system in two variables. Each of these functions is of period 2% both in x and y. The functions of the system (2.1) are orthogonal on the square K (on < x < n, -n ( y < x), as well as on any square of the form (a ( x < a + 2n, b ( y ( b + 2x1). In fact

and similarly

Moreover

j'Jcos mx cos ny) (cos rx cos sy) dy

IC - - cos mx cos rx (j": cos ny cos sy dy dx ) A -

- I, cos mx cos rx d~ Sm cos ny cos sy dy = 0, -IC

if rn # r or n # s. The orthogonality of any pair of dzrerent functions of the system (2.1) is proved similarly. A calculation of the norms gives

11 1 11 = 2%; 11 cos mx 11 = 11 sin mx 11 = Ilcosnyll = Ilsinnyll = . \ /Zx;

11 cos mx me ny 11 = (1 sin mx cos ny 11 = Ilcos mx sin ny 1) = llsin mx sin ny 11 = n.

For the Fourier coefficients of the function f(x, y) defined on K, we obtain

Page 190: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

CHAP. 7

J J ~ ( x . Y ) ~ ~ ~ Y ~ ~ Y AG, = K

llcos nyll2

J-, J- f (x, Y) sin in 'h 4 8 0 = 11 sin mx 11 2

JKJf(x, Y) sin ny dx dy Bh = 11 sin ny 11 2

and similarly,

b, = I. I If(x, y) sin m cos ny dr dy, 7c2 K

1 cmn = - I 1 f ( ~ , y) COS m~ sin ny d~ dy,

z2 K

1 dm = - 1 f(x, y) sin m sin ny dr dy, 7t* K

form = 1,2 ,... andn = 1,2 ,.... Instead of Aoo, one usudly writes 4aoo, and then aoo can be found by

using the first of the formulas (2.2), with m = 0, n = 0. Similarly, if instead 0fAm0, Aon, Brio, and &, one writes h m o , hh, +bmo, and *on, respectively, then %o, son, bmo, and CG, can likewise be found by using the appropriate formulas (2.2).

Page 191: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

With this notation, the Fourier series of f(x, y) can be written in the form OD

f(X, y) - Ln[amn COS m+ COB ny + b, sin mx CoS ny m, n=O + Cmn cos mx sin ny + dm# sin mx sin ny], (2.3)

where ) for m = n = 0, .) for m > O , n = O o r m = O , n > O , 1 for m > 0, n > 0,

and the coefficients am#, bm, c ~ , and dmn are calculated by the formulas (2.2) for m = 0, 1,2 ,... and n = 0, 1,2 ,....

The Fourier series off (x, y) can be written more compactly in the complex form

where

We leave the proof of this result to the reader Ft is recommended to go from (2.4) to (2.3)).

As in Ch. 5, it can be shown that the system (2.1) is complete. This means that

so that

Formula (2.6) expresses Parseval's theorem for the case of two variables. The analogs of all the results implied by the completeness of the trigono-

metric system (proved in Ch. 5 for the case of one variable) remain valid, provided that we suitably modify the way in which the results are stated.

Page 192: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

3. The Integral Formula for the Partial Sums of a Double Trigonometric Fourier Series. A Convergence Criterion

Suppose that we have an expansion of the form (2.3), where it is assumed that f(x, y) is of period 27t both in x and in y. If f(x, y) is defined only on K, we extend it periodically (with respect to x and y) onto the whole xpplane. Now let

m n

snXXI Y) = 2 2 hv[aw,, cos cos vy + brv sin px COS vy woo om0

+ cwv cos (UT sin vy + d,. sin px sin vy].

The quantities sm(x, y) (m = 0, 1,2, . . . ; n = 0,1,2,. . .) are called the partial sums of the double Fourier series. According to (2.2)

Recalling the formula for the sum of cosines (see Ch. 3, Sec. 3), we have

1 sin [(m + *)(s - sin + f.)(? - Y)] & dt. Smn(x9 Y) = ;;i lK.If(', 4 sin 10 - , sin [(t - y)/2]

Setting s - x = u, t - y = v and using the periodicity of the integrand, we obtain

This formula is completely analogous to the corresponding formula (4.1) of Ch. 3, proved for the case of one variable.

The following result can be proved by a method similar to that used in Ch. 3, Secs. 6 and 7:

THEOREM. Let f(x, y) be a continuous f i o n defied on a square K, with bounded partial derivatives a f /ax and a f lay. men, the Fourier series off (x, y) converges to f(x, y) at every interior poiht of K in a neighborhood of which the mixed partial derivative a?f/axay exists. If f(x, y) is of period 2% In x and in y and has continuous partial derivatives af/ax, af/ay, azflaxay, then the Fourier ser;es of f(x, y) converges to f (x, y) everywhere.

For the reader who is not accustomed t~ dealing with huble series, we make the following remark, to avoid confusion: The formula

Page 193: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

f ( ~ , r) = 2 Am[amn COS m . ~ COS ny + b,,,,, sin nX 00s n)r m, n=O

+ Cmn cos mx sin ny + d,,,,, sin rnx sin ny]

means that

or more precisely, that given any E > 0, there exists a number N such that the inequality

holds for m > N, n > N. All that has been said above concerning the square K (-x d x < x, -x < y g x) is also applicable to every square Q ( a < x < a + 2 x , b b x g b + 2 x ) .

Example 1. Expand thefunction f (x, y) = xy for - x < x < x, - x < y < x in double Fourier series* The formulas (2.2) give

a,,,,, = b, = C, = 0,

Using the preceding theorem, we can write Q)

xy = 4 2 (-lIrn+" sin mx sin ny (-x < x < n, -x < y < x)*

m,n=l m

Example 2. Expandthesame fiuiction in thesquare 0 < x < 2rr, 0 < y < 2x. The formulas (2.2) give

aoo = 4x2, with the remaining amn = 0,

4x b0=-- 9 with the remaining bmn = 0, m 4n c b = -. -9 with the remaining cmn = 0, n

Therefore we have

sin mx

m= l m 90 sin ny + sin mx sin ny - 2 n Z - n

mn (0 < x < 2x, 0 < y < 2%). n=1 m,n=l

Page 194: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

CHAP. 7

4. Double Fourier Series for a Function with Different Periods in x and y

A problem which frequently arises in the applications is that of expanding in a double trigonometric series a function f(x, y) defined on a rectangle R (-1 < x < I, -h < y < h), or perhaps a function defined for all x and y, with period 22 in x and period U in y. This problem can be reduced to the problem already considered by makin8 the substitutions u = m/l, o 0' xy/h, since then the function

has period 2n in both u and v. If we have

00

Y(U, V ) - 2 &[amn cos mu cos nv + b, sin mu cos no ms n=O

+ Cmn cos mu sin nv + dmn sin mu sin no],

then, returning to the original variables x and y, we obtain

nmx m y f(xs Y) - l m n [amU. ms - xmx xny

I I) + b, sin T m,n=O COs 7

xmx + Cmn COS - 1 sin 7 + dmn sin - I where

xmx nny .y = ; j"Jf(x, Y ) cos - I cos d* dy,

and so forth. In this case, the complex form of the Fourier series becomes

where

All the results of Secs. 2 and 3 remain valid in the present case, provided we appropriately change the way in which the results are stated.

Page 195: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

5. The Fourier Integral as a Limiting Case of the Fourier Series

Let f(x) be a function defined for all real x, and let f(x) be piecewise smooth (~~,ntinuous or discontinuous) on every finite interval [ - I , 4 Then on every such interval, f(x) can be expanded as a Fourier series

where

(see Ch. 1, SCC. 15). [At points of discontinuity, we have to write +V(r + 0) + f(x - 0) J instead of f (x) in (5. I).] Substituting the expressions for a,, and bn in (Sol), we obtain

We now assume that Ax) is absolutely integrable on the whole x-axis, i.., we suppose that the integral

exists. Then as I -, co (x fixed), (5.2) becomes

To investigate the limit on the r i a we set

X 2X nrt A, = T XI = T..., A,, =- I '..*'

af€er which the sum in (5.3) takes the form

This reminds one of the sum defining the integral of the function

Page 196: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

of the variable A over the interval [0, +a ] . Therefore, it is natural to expect that as I+ oo, (5.4) goes into an improper double integral, i.e., it can be anticipated that we have the formula

f(x) = ir dAr /(U)cos?t(u - X) du. X 0 -00

[instead of (5.3)]. Of course, this argument is not rigorous, but at least we now know what formula to expect. It will be proved below that (5.5) is actually valid, with our assumptions concerning f (x) (and even with somewhat weaker assumptions). In this regard, we point out once more that in (5.5) we have to write +V(x + 0) + f(x - O)] instead of f (x) at points where f (x) is discontinuous.

The inner integral in (5.5) is called the Fourier integral, and the entire formula is called the Fourier integral theorem. Using the formula for the cosine of a difference, we can write

f(x) = & [ 4 ~ ) cos AX + HA) sin AX] dA (5.6)

instead of (5.5), where 1 a 1

a@) = n loo ~ ( u ) cos AU ~ii(, HA) = - 10 f(u) sin AU ttu. (5.7) X -a0

The reader will immediately notice a resemblance between (5.6) and a Fourier series: The sum has been replaced by an integral and the formula involves a continuously varying parameter A instead of the integral parameter n. Moreover, the coefficients a(A) and HA) are quite like Fourier coefficients.

6. Improper Integrals Depending on a Parameter

Consider the integral

and suppose that it is convergent for a < A < p. We shall say that it is un~ormly convergent for a < A < if for every r > 0, there exists a number L such that

for all I L and for all A (a ( A $ p). We begin by proving the following result: A necessary and suflcient condition for the integral (6.1) to be uniformly

convergent is that for every sequence of numbers x0 = a < x1 < x z . < - * - < x,, <-=,

lim x, = a,

Page 197: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

the series

whose terms are functions of A, should be unifrmly convergent for a c A < f3. The proof goes as follows: If (6.2) holds, then for x,, 3 L we have

which means that the series (6.4) is uniformly convergent. Conversely, suppose that the series (6.4) is uniformly convergent for every Bequence of the form (6.3). Then, if the integral (6.1) were not uniformly convergent, this would imply the existence for at least one o > 0 of arbitrarily large numbers xl < x2 < < xn < ., such that

for each q (i = 1,2, . . .) and some A. But this is impossible, since if we choose these values of x, as the sequence (6.3), then (6.5) must hold for sufficiently large n. This contradiction shows that the integral (6.1) is uniformly convergent.

This result implies

THEOREM 1. If F(x, A) is continuous, regarded as a function of two variables (or if F(x, 1) has discontinuities at a finite number of values of x in each jinite interval, while remaining integrable with respect to x and continuous in A) and if the integral (6.1) is unvormly convergent for a d X ( f3, then the integral is a continu011~fUnction of A.

Proof. Every term of the series (6.4) is a continuous function of h (by a property of integrals with finite limits), and since this series converges uniformly, its sum, i.e., the integral (6.1), is a continuous function.

T H ~ M 2. With the hypotheses of Theorem 1, we have

Page 198: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

1 84 DOUBLE FOURIER SBRIES. THE FOURIER INTEGRAL CHAP. 7

Proof. Because of the uniform convergence of the series (6.4), it can be integrated term by term, and the order of integration can be interchanged in every term (this is legitimate forjhite intervals). Thus, we obtain

THEOREM 3. Suppose that F(x, A) b contfnuow, regardcd as ajknctbn of two varhbles, and has a continuotu partial derivative aF(x, A)/aX. Suppose further that the integrak

exkt and that the second of them 25 uniformly convergent for a 6 h 6 $. Zkn we have

PrwJ The series

whose terms are the derivatives of the corresponding terms of the series (6.4) (since forfInfte intervals, differentiation under the sign of integra- tion is legitimate), is uniformly convergent. But then we are j u a e d in differentiating (6.4) term by term (see Theorem 2 of Ch. 1, Sec. 4).

The next theorem gives a very useful criterion for the uniform convergence of the integral (6.1) :

and the integral

exists, then the iittegral(6.1) b w w y convergent. (The requirements on F(x, h), regarded as a function of x, are the same as in Theorem 1.)

Page 199: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Proof. The absolute values of the terms of the series (6.4) do not exceed the terms of the convergent numerical series

which proves the theorem.

Instead of (6.1), we can consider integrals of the form

with the condition (6.2) replaced by the condition

for the first integral, and by the two conditions

for the semnd integral. The theorems just proved remain valid for integrals of this form. In fact, the proofs reduce to the case already considered if we make the substitution x = - y in the first integral, and divide the second integral into the two integrals

7. Two Lemmas

We now refine the results of Ch. 3, Sec. 2:

LEMMA 1 . If f(x) is absolutely integrable on a < x < a, then

lim f(u) sin lu du = 0 C-*m a

(7.1)

(where I takes arbitrary values).

Proof. For any given c > 0

( r f(u) sin d u G

for sufficiently large b. Moreover, by the theorem of Ch. 3, Sec. 2

I Jbf(u) sin zu dul G I a

Page 200: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

CHAP. 7

for sufficiently large 2. Therefore

(r f(u) sin lu du( P a

for all sufficiently large I, which proves (7.1).

Remurk. Instead of the integral from a to 00, we can consider the integrals from -00 to a and from -a, to a. Also, instead of sin Zu, we can consider cos Zu. The proofs are essentially the same.

LEMMA 2. If the function f(x) b absolutely integrable on the whole x-axis, and f (x) has ZeB-hand and right-hand limits at the point x, then

sin lu lim !f'O JCX + u)- du = f (x+O)+f(x-O) . (7.2) b o o -00 24 2

Proof. Given any o > 0, the inequality

lp 'It -8 l f (x+ * ) p i 4 <;

holds for sufficiently small 8 > 0. The function (l/u)f(x + u) is absolutely integrable for -oo < u < 8 and 8 < u < a. Therefore, by Lemma 1

sin lu 1 - 8 U hao % -00

li( du = 0. (7.3) ~ h n i I Q ~ + u ) - ~ = l i m - j f(x+u)- l+oo 'It 8 U

Now consider the equality

sin mu $ f(' + U ) 2 sh (@) c h r = f(x + 0) + f ( x - 0)

2 ¶

m-boon -n

proved in Ch. 3, Sec. 7, where m = n + f for integral n. This can be written as

1 8 sin mu '"O - I f(% + 2 sin du= f(x + 0) + f(x - 01, (7.4) m-boos -8 2

since the integrals on the intervals [ -7 i , -81, [6, X ] go to zero as m + 00, because of the absolute integrability of the function

f (x + u) 2 sin (2412)

on these intervals (see Q 3, Sec. 2). Next, we observe that the integral in the left-hand side of (7.4) differs

from the integral ' f(x + u) sin mu

24 du

X -8

Page 201: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 7

by the quantity

I - A] sin mu b, I8 f(x + [2 sin (uj2) u X -8

where the function in brackets is continuous, if it is regarded as being equal to zero at the origin. (This can be seen by applying L'Hospital's rule.) But by Ch. 3, Sec. 2, the integral (7.6) converges to zero as m + 00. Therefore, instead of (7.4), we can write

1 s lim - f(x + u)

sin mu du= f (x + 0) +f(x - 0)

m+ooZ -8 U 2 (7.7)

Now let m < I < m + 1, so that I = m + 0, where 0 < 0 < 1. Applying the mean value theorem, we obtain

sin lu - sin mu = (I - m) cos hu = 0 cos hu,

U

where h lies between m and I. Therefore

sin lu 1 8 - - I-&f(x + u) sin mu

St St U

for any I. If I is large, then m is also large, and therefore by (7.3, we obtain

I sin mu s 2 u d u < 2

fgr the values of m corresponding to large values of 1. Together witk (M), this inequality gives

I sin lu 2 -8 U

far all sufficiently large I. By (7.3), instead of (7.9), we can write

I f ( ~ + O ) + f ( x - 0 ) - - 1 00 sin lu 2 du < e, n.I-OOf(x + ~1-i- I

for sufficiently large I. This proves (7.2).

Page 202: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

CHAP. 7

8. Proof of the Fourier Integral Theorem

Suppose that f(x) is absolutely integrable on the whole x-axis. Then, by the very definition of an improper integral, we have

i.e., the existence of the integral in the left-hand side of (8.1) is equivalent to the existence of the limit in the right-hand side. But ths integral

is uniformly convergent for -a, < A < oo, since

and f(u) is absolutely integrable on the whole x-axis (see Theorem 4 of Sec. 6). It follows by Theorem 2 of Sec. 6 that

= J",f(.> sin l(u - sin lu

U - X x)du = IP -00 f (x + u)- U du,

where we have first made the substitution u - x = 0 and then changed v back to u. By (8.1), we have

1 sin lu iJ%p f(u)aw~(u - x ) h = lim -10 f(x + +I- u h. 7C 0 -a0 I+ao x -a

Now, if the function f(x) has a left-hand and a right-hand derivative at the point x, then by Lemma 2 of Sec. 7, the limit in (8.2) exists and is equal to +V(x + 0) + f(x - O)]. Therefore, the integral in the left-hand side of (8.2) exists and we have

At the continuity points of f(x), the right-hand side of (8.3) coincides with f(x). Thus, we have the following results:

1. Iff (x) i& absolutely integrable on the whole x-axis, the Fourier integral theorem holh at every point where f(x) has both a lep-hand and a right-hand derhatwe.

2. vf(x) is absolutely integrable on the whole x-axis and iff(%) &piecewise smooth b7, every flnite internal, then the Fourier integral theorem holdr for d r

Page 203: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

9. Different Forms of the Fourier Integral Theorem

Assume that f(x) is absolutely integrable on the whole x-axis and consider the integral

which converges uniformly for - 00 < A < GO, since

(see Theorem 4 of Sec. 6). Therefore the integral represents a continuous function of A, which is obviously odd. But then2

1' d~ JIOD f(u) sin ~ ( u - x) du I+m -1 -a0

= sx I" M U ) sin ~ ( u - x) ~ h r 0. -a0 --a0

On the other hand, the integral

00

L a f(u) cos q u - x) m(

represents an even function of A. Therefore, instead of (5.9, we can write

f(x) = I &Js f(u)[ow ~u - x) + isin ~ ( u - 2z -a0 --QO

(9.1) = I I" & r f (u)e'X(&x) du.

27t -a, -4

which is the complex form of the Fourier integral theorem. Next, we write (5.5) in the form

f(x) = cos AX (I: f(u) cos b du) dA 7r 0

(9.2) + !. sin AX ( / l f ( u ) sin AU tiu) A. 7t 0

2 If the integral with respect to X in the right-hand side of this formula does not exist in the usuaf sense, then we interpret it as

a limit which in this case obviously exists and equals 0 (the Cauchy prfncIpa1 due of the integral).

Page 204: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

CHAP. 7

If f(x) is even, then a0

I, f(u) cos hu L = 2 j- f(u) cos Xu du, 0

a0 j'*f(u) sin L = 0,

and (9.2) becomes

f(x) = z J'O ms hr (C f(u) cos hu du) a. X 0

Similarly, iff (x) is odd, we obtain

f(x) = -Z sin AX (l:f(u) sin ?a du) a. = 0

If f(x) is defined only on [0, a], then (9.3) gives the even extension of f (x) onto the whole x-axis, while (9.4) gives the odd extemion of f(x). Thus, both formulas are applicable for positive x, but for negative x they give different values of f(x). We note that if f(x) is continuous, then (9.3) is always valid at x = 0, whereas (9.4) is valid only if f(0) = 0. w s is because when x = 0, the integral (9.4) takes the value +[f(+O) + f(-O)b and this value is always zero for the odd extension off (x).]

* 10. The Fourier Transform

Given the function f (x), the new function

is called the Fourier transform of f(x). If the Fourier integral theorem holds for f(x), then by (9.1) we have

i.e., f (x) is the (inverse) Fourier transform of F(x). Thus, the function (10.1) can be regarded as the solution of the integral equation (10.2), when f(x) is a given function.

We now note some properties of the Fourier transform (10.1).

1. Iff (x) fi abso2utely integrable on the whole x-axis, then the fiulction fix) is continuous for a21 x and convergw to zero g 1x1 -c a.

Proof. The continuity of F(x) follows from the uniform convergence (in x) of the integral (10. I), since

Ic'"l=l. If(u)e'-l=If(~)I,

Page 205: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 10 DOUBLE FOURIER SERIES. THE FOURIER INTEGRAL 1 9 1

and the integral

exists. (All the considerations of Sec. 6 continue to hold in the case where f(x) is a complex-valued function.) Moreover

lim F(x) Ixl+-

1 a0 = - [ lim J" f(u) cos xu du + i lim f(u) sin xu h] = 0,

2 Ixl+ol -00 1x14- 1-rn by the remark to Lemma 1 of Sec. 7.

2. If the fwction ~ l f (x) [J absolutely integrable on the whole x-axis (n is a positive integer), then F(x) is dffrrentiable n times, where

F(k)(x) = - f (u)ule'm du (k = 1,2, . . . , n), (10.3)

and all these derivatives converge to zero as 1x1 -t co.

Proof. The formula (10.3) can be obtained by differentiating (10.1) behind the integral sign, since each time we obtain an integral which converges uniformly in x. This follows from the relations

where the functions on the right are absolutely integrable. (See Theorem 3 of Sec. 6.) To prove that the derivatives flk)(x) converge to zero as 1x1 + co, we again use the remark to Lemma 1 of Sec. 7.

3. Iffx) is continuous and converges to zero as 1x1 + co and gf(x) is absolutely integrable on the whole x-axis, then

4. f (x) is absolutely integrable on the whole x-axis and [f

ac 1x1 + a, then

1 i j" (l:f ( t ) dt) e" du = - Zy"'. 7t --a, X

To prove the last two formulas, we use integration by parts. These fohulas show that differentiating the original function f(x) corresponds to

Page 206: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

192 DO- POURSER SERIES. THE FOURIER INTEGRAL CHAP. 7

multiplying its Fourier transform F(x) by x/i, while integrating f(x) com- sponds to dividing F(x) by x/i. This idea of reducing complicated mathe- nutical operations on the original function to simple algebraic operations on its transform (and then taking the inverse transform of the final result) is the basis for the operational caldw, a very important branch of applied mathematics.

Next, we consider transforms of a somewhat different form. The function

is called the (Fourier) cosine truqfbrrn of the function f(x). If the Fourier integral theomm holds for f(x), then it follows from (9.3) that

i.e., f(x) is i e l f the Fourier cosine transform of F@). In other words, the functions f and F are cosine transforms of each other. Similarly, the hnction

@(A) = f(u) sin hu du s o

is called the (Fourier) sine trmfom of f(x), and (9.4) gives

i.e., just as in the case of cosine transforms, f and cP are sine transforms of each other.

The function (10.4) can be regarded as the solution of the integral equation (10.5) [where f (x) is a given function], and the function (10.6) can be regarded as the solution of the integral equation (10.7).

We now illustrate the use of Fourier cosine and sine transforms by evaluating some integrals.

Example 1. Letf(x) = gu (a > 0, x 3 0). This function is integrable for 0 C x < ao and has a derivative everywhere. Integrating by parts, we find

Page 207: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Then (10.5) and (10.7) give

Example 2. If 1 for 0 d x < a,

for x = a, 0 for x > a,

then obviously - -

2 sin aA = J:r- X 0 h.1- JET,

and by (10.5)

a sin ah cos X A 6h 1 for 0 < x < a, f(x) = 21 A for x = a, x 0 0 for x > a.

In particular9 for x = a

1 - = 11 a sin 2 a ~ 2 x o h fi

and setting a = +, we obtain

* I I. The Spectral Function

It is easy to see that equation (9.1) can be written in the form

since dwx-u) = cos h(x - u) + i sin h(x - u)

= 00s l(u - X) -- i sin h(x - u),

and since the integral containing the sine vanishes (see Sec. 9). Now we set3

3 Sometimes the factor 1/2sr is omitted in the formula for A Q ; of course, the factor then reappear8 in (1 1.3).

Page 208: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

194 DOUBLE FOURIER SERIBS. THB FOURIER INTEGRAL CHAP. 7

This function (which is in general complex) is of great importance in electrical engineering, and is called the spectral function (synonymously, the spectral demity or spectm) of f(x). According to (1 1.1) and (1 1.2)

(1 1.2) is the analog of formula (14.7) of Ch 1 (which gives the values of the complex Favler coeflclents), and (11.3) is the analog of formula (14.6) of Ch. 1.

Exampla Find the spectral function of the function

1 for 1x1 < a, 0 for 1x1 z a,

shown in Fig. 42.

According to (1 1.2)

1 eth - e - I & 1 sin ax =- c c--,

27t ih 7t h

and therefore in this case A@) turns out to be real. The graph of A Q is shown in Fig. 43 (for a = 1).

Page 209: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

PROBLEMS DOUBLE FOURIER SERIES. THE FOURIER INTEGRAL 195

PROBLEMS

1. Show that if f(x, y) = g(x)lr0), where

then

2. Show that iff (x, y) = g(x + y), where

then

3. Expand the following functions in double Fourier series in the region - IcQx,yQx:

a) f ( x , u ) = x + y ; b) f(x, Y) = (x + yl2 ; C) f (~ , Y) = Isin (x + y)l

4. Find the Fourier sine transform (10.6) of each of the following functions: sinx for 0 6 x < rc, a) f ( x ) = e - x m r x ( ~ ~ x < a ) ; b) j ~ ( x ) = { ~ for x > sr;

for O < x < 1, d) f(x) = xe-X (0 < x < a). for x > l ;

5. Find the Fourier cosine transform (10.4) of each of the functions of the preceding problem.

6. Solve the following integral equations for f(x) (0 Q x < a):

a) f (x) sin lx abr = g(X), where g(A) = 1 for 0 < h < rc,

for X > x ;

b) JOaf(x) cos XY d~ = e-a;

C) & f(x) sin d~ = *-.A.

7. Let f (x) and g(x) be absolutely integrable on - a < x < a, and let

Let F(h), G(X) and H(X) be the Fourier transforms [see equation (10.1)] of f (x), g(x) and h(x), respectively. Show that

Page 210: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

CHAP* 7

a0 L, f ( x - Y ) ~ ( Y ) di. = j- -00 g(x - y)fW v)* The function h is called the convolution off and g, written h = f *g.

80 Find h = f *g (cf. preceding problem) where x for O < x < l ,

a) f(x) = rx, = { 0 for x > 1; b) f (x) = g(x) = e-X;

1 for 0 4 x < 1, for O < x < 1, for x > l ; for x > 1;

x for 0 6 x 4 , 0 for x > l .

(In every case, f (x) = g(x) = 0 for x < 0.)

9. Solve the following integral equations : for x < 0,

a) f - Y for x a o ;

Page 211: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

BESSEL FUNCTIONS AND FOURIER-BESSEL SERIES

I. Bessel's Equation

By BesseI's equation is meant the second order differential equation

x2y" + xy' + (x2 - p2)y = 0, (1.1)

or equivalently

where p is a constant [which we call the order of (1. I)], and the prime denotes differentiation with respect to x. With the exception of certain special values of p, the solution of (1.1) cannot be expressed in terms of elementary functions (in finite form); this leads to the socalled Bessel fmctlo~~s, which have many applications in engineering and physics. Tables of Bessel functions are available for use in practical calculations.

Since Bessel's equation is linear, it has a general solution of the form

where yl, yz are any two linearly independent particular solutions of Bessel's equation, and Cl, C2 are arbitrary constants. Thus, to find the general solution of (1.1). it is sufficient to find any two linearly independent solutions of (1 . 1).

197

Page 212: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

2. Bessel Functions of the First Kind of Nonnegative Order

Suppose that p 2 0. To simplify subsequent calculations, we make the substitution

y = xpz

in the equation (1.1). Since

the function z satisfies the equation

We now look for a solution of this equation in the form of a power series

z = c(,+ c1x+c2x2+***+ c , p J + e * . .

Elementary calculations give

z' = ct + 2c2x + 3c3x2 + 4 ~ ~ 3 + **+ (n + ~)c,~P+' + 9

Z" = 2% + 2 . 3 ~ ~ ~ + 3 . 4 ~ 4 ~ ~ + * = * + (n + l)(n + 2)cn+ze + s o * .

Substituting these series in (2.2), we obtain

If this equation is to be satisfied, all the coefficients of the different powers of x must vanish, i.e.,

Therefore, we have

Page 213: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 2 B ~ E L FUNCTIONS AND FOURIER-BESSEL SERIES 1 99

It follows from (2.3) and (2.4) that

and in general

Thus, the series

where co is an arbitrary constant, gives a formal solution of the equation (2.2). By using the familiar ratio test, it is easy to show that this series converges for all x. Since term by term differentiation of a power series is always legitimate (inside the interval of convergence), z is actually a solution of (2.2). But then, the function

where co is arbitrary, is a solution of (1.1). It is customary to set

where I' is the well-known gamma function of mathematical analysis, which has the following properties :

1) r( i ) = 1, 2) l'(p + 1) = PI'@) for any p, 3) l'(p + 1) = p! for positive integral p.

(We shall discuss the gamma function in more detail in the next section.)

Page 214: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

200 PUNCIIONS AND F o U R m R - m SBRtBS CHAP. 8

If the constant q is defined by (2.a then the series (2S) gives the Bessel j k t h of the f i s t k i d of orhr p (for the time being, p ) 0), denoted by J',x):

But according to the properties of the gamma function, we have

(P + + 2). *(p + m)r(p + 1) = (p + 2)b + 3). . .(P + + 2)

=(P + 3)b + 4).**@+ m)r(p + 3)

@ + m)r@ + m) = r(p + m + I), and therefore

(see also Property 1 of the gamma fbnction). In particular, for p = 0

wherewemustset01 = 1. Forp = 1

In general, for positive integral p

(- l)n(x/2)ZHJ'* m=O m!(p + m)!

Formulas (2.9) and (2.10) show that if p = 0 or if p is an even integer, then JJx) is an even function (since it contains only even powers of x). On

Page 215: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 2 BESsEL FUNCTIONS AND FOURIER-BESSEL SERIES 20 1

the other hand, i f p is an odd integer, then Jp(x) is an odd function (since it contains only odd powers of x)- The graphs of the functions y = Jo(x) and y = Jl(x) are shown in Fig. 44.

Remark. It should be noted that if x < 0 and i f p is not an integer, then in general Jp(x) will take complex values [see (2.7)]. In order to avoid complex values, we shall consider Jp(x) only for x 3 0 (when p is not an integer).

3. The Gamma Function

For p > 0, the gamma function is usually defined by the formula

(This improper integral has meaning only for p > 0.) We now verify that (3.1) actually has the properties given in Sec. 2.

Integrating by parts, we obtain

The first term on the right vanishes (provided, of course, that p > 0), and the integral is just r@). This proves Property 2.

3) If p is a positive integer, then using Property 2 we obtain

or, in view of Property 1

r(p + 1) = p !

Thus, for p > 0, the function defined by (3.1) actually has the required properties.

Page 216: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

202 BmEL FUNCTIONS AND FOURIER-BESSEL SERIES CHAP. 8

To extend the function r(p) to all real values of p, we start with the formula

If - 1 < p < 0, the right-hand side of (3.2) has meaning, since then 0 < p + 1 < 1. Therefore, (3.2) can be used to define r(p) for - 1 < p < 0. Incidentally, we note that as p + 0, the numerator in the right-hand side of (3.2) approaches 1 while the denominator approaches 0, and hence we have

NOW let -2 < p < - 1. Then - 1 < p + 1 < 0, and the right-hand side of (3.2) again has meaning. Therefore, (3.2) can be used to define r@) for - 2 < p < - 1. If p + - 1, then it follows from (3.2) that r(p) + a, and hence we have

Next, we consider the values - 3 < p < - 2, and so forth. Thus, step by step, we can define r(p) for all negative values of p, with r(p) = co for p = 0, - 1, - 2, . . . . In other words, (3.1) can be used to define r@) for p > 0 and then (3.2) can be used to define I?@) for all realp. Moreover, by its very construction, r(p) has the three properties listed above.

4. Bessel Functions of the First Kind of Negative Order

Since p2 appears in equation (1.1), it is natural to expect that the con- siderations of Sec. 2 will be applicable to - p as well asp and will also lead to a solution of (1.1). Replacing p by - p in (2.8) gives

We note that for integral p and m = 0, 1,2, . . . , p - 1, the quantity -p + m + 1 takes negative values and the value zero. Therefore, r(-p + m + 1) = ao for these values of m, and the corresponding terms in the series (4.1) are taken to be zero. Thus, for integral p

Page 217: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 4 BESS~L PUNCTIONS AND F O ~ - B E S S ~ L SER~ES 203

or, ifwe set m = p + k

If p is not an integer, then none of the denominators in (4.1) becomes i w t e . The ratio test shows that the series (4.1) converges for all x # 0 if p is not an integer and for all x if p is an integer [see (4.2) 1.

The function LP(x) is again called a Besselfunction of the Prst kind, tbis time of order -p . It is easily verified by direct substitution that LP(x) is actually a solution of (1.1). We leave this verification to the reader.1

For what follows, it is useful to observe that the formulas (2.8) and (4.1) can be combined in one formula

where the number p can be either positive or negative. The remark made at the end of Sec. 2 applies to the case of fractional values of p of either sign.

5. The General Solution of Bessel's Equation

Consider first the case where the number p > 0 is not an integer. Then, the functions J b ) and J-p(x) cannot be linearly dependent, i.e., there cannot exist a constant C such that

To see this, we observe that for x = 0, the function Jp(x) vanishes, whereas J-Jx) becomes infinite [see (2.8) and (4. I)]. Thus, if p is not an integer, the general solution of Bessel's equation (1.1) has the form

where Cl and C2 are arbitrary constants [cf. (1.211. If p 3 0 is an integer, then by (4.2) the functions J,(X) and J-(x) are

linearly dependent, and hence in this case, (5.1) does not give the general solution. Therefore, for integral p we have to find another particular solution of (1.1), which is linearly independent of Jp(x). This particular solution Yp(x) is the so-called Bessel f w o n of the second kind, to be

1 It should be noted that for some values ofp (e.g., for integralp), part of the argument given in Sec. 2 ceases to be legitimate when applied to -p, since it leads to fractions with zero denominators [see e.g. (2.4)]. Nevertheless, the final formula (2.8) has meaning when p is changed to -p, and in fact gives a solution of (1.1), as noted.

Page 218: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

204 BBSSBL FUNCTIONS AND FOURIER-BESSEL SERIES CHAP. 8

discussed in the next section. Thus, when p is an integer, the general solution of (1.1) has the form

Y = ClJp(~1 + C2Yp(x).

6. Bessel Functions of the Second Kind

For fmctional p, the Beswl function of the second kind is obtained from (5.1) by a special choice of the constants Cl and Cz, i.e., we set

YJx) = $Ax) 00t prr - LP(x) csc prr = Jp(x) cos - J-dx)e (6.1) sin pn

For integral p, (6.1) is indeterminate; in fact, the numerator reduces to (- l)pJ',(x) - J&(x) which vanishes according to (4.2), and the denominator also vanishes. This suggests the following question: Can the indeterminacy be "removed" by taking the limit of the ratio (6.1) as p approaches integral values, and will this limit give the required solution for integral p? As we now show, the answer is in the ahnative.

According to L'Hospital's rule

Yn(x) = lim (slap) [ JAx) cos F - J-Ax)] ~n (WP) sin F

= lim cos ~rr(aIa~)J;(x) - -A~) sin ~n - (a/B~V-p(x) IHn 7c cos prr

We substitute the series (2.8) and (4.1) in this expression, differentiate them with respect to p, and then set the arbitrary index p equal to the integral index n. After some manipulations, the details of which we omit (since they involve spedal properties of the gamma function and are rather tedious), we obtain

where C = 0.577215664901 532 is the so-called Euler's constant. In particular, when n = 0

Page 219: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Substitution of the function Yn(x) into the equation (1.1) with p = n shows that Ym(x) is actually a solution of (1.1). Moreover, the fmctions Jn(x) and Yn(x) cannot be linearly dependent, since for x = 0, Jn(x) has a fmite value, whereas Ydx) becomes infinite. Therefore, Y&) is the re- quired particular solution of (1.1) (cf. the end of Sec. 5). The graph of the function y = Ydx) is shown in Fig. 45.

7. Relations between Bessel Functions of Different Orders

For any p, we have the formulas

Similar formulas hold for the corresponding Bessel functions of the second kind.

Proof. By (4.3), we have

for any p, which proves (7.1). Formula (7.2) is proved in the same way. To prove the corresponding formulas for Bessel functions of the second kind, we first replace p by -p, obtaining

Page 220: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

206 BESSEL FUNCTIONS AND FOURIER-BESSEL SERIES CHAP. 8

Assuming that p is not an integer, we multiply (7.1) by cot prr and subtract (7.3) multiplied by cscpn. The result is

A [xp (x) cos prr + J-*i(x) dx sin prr sin pn

since

cos@ - l)n = - cosprr, sin@ - 1)n = - s i n p .

Thus, for fmtional p

(6.l)l. Next we replace p by -p in (7.1), obtaining

Then we multiply (7.2) by cotprr and subtract (7.5) multiplied by cscprr. The result is

JP(x) ~ 0 s F - J-p(x)] sin pie

= X - P - J*I(X) P1E - J-p-l(~) sin prc

since

cos ( p + I)= = - cos prs, sin ( p + 1)n = - sin par.

Therefore, for fractional p [see (6.111

The formulas (7.5) and (7.6) are the analogs of (7.1) and (7.2), and also hold for integral p (as can be seen by letting p approach integral values).

From (7.1) and (7.21, we obtain the formulas

Page 221: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 7

and similar formulas for Bessel functions of the second kind. In fabt, it follows from (7.1) that

xpJtP(x) + pxplJp(x) = xpJ P-1 (x),

from which (7.7) is obtained by dividing by xp-1. In the same way, we obtain (7.8) from (7.2). To obtain (7.9), we add (7.7) and (7.8) and divide the result by x. Finally, (7.10) is obtained from (7.7) and (7.8) by sub- traction and subsequent division by r

The utility of these formulas will be repeatedly demonstrated below, and hence we shall not discuss them now. However, it should be noted that (7.10) shows how to calculate the values of the function Jp(x) for any positive or negative integral p from a knowledge of the functions Jo(x) and Jl(x). A similar remark applies to Bessel functions of the second kind.

8. Bessel Functions of the First Kind of Half-Integral Order

Referring to (2.7), we first consider

- - 1

R*) sin x.

According to (3.1)

1 1 ( i ) = 2r(Z) = :I,.-~- dx = $' rta dt, <x 0

where the integral equals fi12.2 Therefore

Jllz(x) = d2/7tx sin x,

2 To see this, write

Page 222: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

208 B m FUNCTIONS AND FOUIUBR-BESSEL SERIBS CHAP. 8

and similarly, we find

JOlJ2(x) = 4- ws X.

Thus, the functions JlJ2(x) and LlJ2(x) can be expressed in terms of elementary function8. But then it is an immediate consequence of formula (7.10) that any function JJx) where p = n + f. and n is integral, can be e x p r d in terms of elementary functions. For example, setting p = in (7.10), we obtain

whence

1 - J3&) = ;Z JlJ2(x) - J-~ J2(x) = 1 2 1 ~ ~ 3 sin x - 1 / 2 2 cos x.

Similarly, settin8 p = in (7.10), we can 5 d JSJz(x), etc.

9. Asymptotic Formulas for the Bessel Functions

In this d o n , we derive formulas that permit us to readily determine the behavior of Bessel functions for large values of x (so-called asymptotic formulas). First of all, we transform equation (1.1) by making the sub- stitution

The resulting equation for the fundion z is

If we set

then (9.2) becomes

2' + (1- + p)z = 0.

For lare x, the function p = p(x) becomes very small. Therefore, it is natural to expect that for large x, the solution of the equation (9.4) will not differ much from the solution of the equation

Page 223: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 9 BESSEL ~MCTIONS AND FOURIER-BESSBL 209

i.e., from the function

z = A sin (x + o) (A = const, o = const).

Now, let z be a solution of (9.4) which is not identically zero. Then, in view of what has just been said, it is natural to assume that there exist functions a = a(x) and 8 = 8(x) such that

z = a sin (x + 8). (9.5)

where a and 6 converge to definite finite limits as x+ a. To prove the existence of such functions, we consider the equation

as well as (9.5). and we then regard (9.5) and (9.6) as a system of equations where the left-hand sides are known and the functions a and 6 are unknowns.

It follows from (9.4) and (9.5) that

z" = - (1 + p)a sin (x + a),

and from (9.6) that

z" = a' cos (x + 8) - a(l + 8') sin (X + 8).

which imply at once that

Differentiating (9.5) gives

i = a' sin (x + 8) + a(1 + 83 cos (x + 8).

Comparing this with (9.6), we easily obtain

as' tan (x + 8) = - -. a'

Multiplying (9.7) and (9.8), we find

whence

6' = p sin2 (x + 6).

Then (9.8) implies

a' - = - 8' tan (x + 6)

= - p sin (x + 6) cos (x + 6). a

We observe that the denominator a of the ratio a'/a cannot vanish. In fact, if a vanishes, then it follows from (9.5) and (9.6) that z and z' vanish

Page 224: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

2 I O BEssEL FUNmONS AND FOURIER-BESSEL SERIES CHAP. 8

simultaneously, i.e., the solution z satisfies zero initial conditions at some point x = xo. However, because of the uniqueness of a solution satisfying given initial conditions, this implies that z is identically zero, contrary to hypothesis.

The required function 6 is found from the differential equation (9.9), and the initial condition for 6 can be found from the initial conditions for z by using (9.5) and (9.6) and eliminating a (e.g., by division). Then, from a knowledge of 6, we can easily find a from (9. lo), and the initial condition for a is obtained from the initial conditions for z and 6 by using (9.5) and (9.6).

Thus, all that remains is to analyze the asymptotic behavior of a and 6. Since

it follows from (9.3) and (9.9) that

If we take the limit as b+ a, the resulting improper integral obviously converges (since the integrand does not exceed lltz). Therefore, the limit of 6(b) as b + a, also exists. If we set

lim 6(b) = a, b o o

then

But

and hence

a sin2 (t + 6) o < m x l % p dt < m.

In other words, the function

I.. sin2 (t + 6) dt q(x) = - mx X t2

is bounded for any x, and (9.1 1) can be written in the form

Page 225: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 9

Moreover

BEssEt FUNCTIONS AND FOURIER-BESSEL SERIES 2 1 1

a' -- = (In a)', a

and hence

In a(#) = In a(b) + in I sin (t + 6) cos (t + 6) X t2

Just as before, we pass to the limit as b+ a, and note that the resulting improper integral converges. This implies the existence of a finite limit as b + a, for ln a(b) and hence for a@). We set

lim a@) = A, b+ao

where A # 0, since otherwise in a(b) + m as b + a,. Then we have

lna(x) = In A + nr .. sin (t + 6) cos ( t + 8) t2

dt. X

Just as we proved the boundedness of the function q(x), we can also prove the boundedness of the function

a sin (t + 6) cos (t + 6) dt. d.1 = mx J X t2

Then

whence

= A exp k(x)/xl*

According to Taylor's theorem, for any t we have

e t = l + t @ t ( 0 < 8 < 1).

and hence, writing t = p(x)/x, we obtain

exp [~p(x) /x l . exp k(x)/xl = 1 + -;;-

Page 226: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

2 12 BBSSBL FUNCTIONS AND FOURIER-BESSEL SERIES CHAP. 8

The function exp [@(x)/x] is obviously bounded as x + ao. Therefore, we can write

where e(x) remains bounded as x+ a. The formulas (9.12) and (9.13) confirm our conjecture about the way the

functions a and 6 behave as x+ do and about the behavior of the solution z of the equation (9.4). [Cf. (9.9.1 Substituting (9.12) and (9.13) into (9.5) gives

Next, we transform the last factor in (9.14). According to Taylor's theorem

sin (a + t) = sin a + t cos (a + 8t) (0 < 8 < 1).

Setting a = x + a, t = q(x)/x, we obtain

sin x + o + ( C(x) *I) = sin (X + a ) + 7 X

where

is a function which is bounded for all x. Therefore, it follows from (9.14) that

= A sin (x + a) + A Ux) sin (x + a ) + (1 + e(x)/x)c(x) X

where r(x) is bounded as x-, co. Thus, we have found an asymptotic formula for the solution of (9.2).

Page 227: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 9 BWSEL FUNCTIONS AND FOURIER-BESSBL =RIBS 2 1 3

To go to the case of Bessel's equation, we use (9.1) and obtain

A r(x) = - sin (x + o) + - fi d x '

where A = const, o = const and r(x) is bounded as x+ ao. This formula shows that for large x, any solution of Bessel's equation differs very little from the damped sinusoid

A y = sin (x + a).

In particular, the above considerations apply to the Bessel functions Jp(x) and Yp(x). A more exact calculation, which we omit, shows that

psc Jp(x) = sin (x - +

Yp(x) = - sin

where the functions rp(x) and pp(x) remain bounded as x-c a. For subsequent purposes, besides the formula (gels), it is useful to have

the comsponding formula for z'. Therefore, we substitute the expressions (9.12) and (9.13) for 6 and a into (9.6). The result is

If we transform this expression in the same way as we transformed (9.14) [which led to (9.1511, we obtain

where the function s(x) is bounded as x+ 00.

10. Zeros of Bessel Functions and Related Functions

It follows at once from formula (9.15) that any solution of Bessel's equation has an infinite number of positive zeros and that these atras are close to the zeros of the function sin (x + a), i.e., to the numbers of the form

where n is an integer. We now show that for sufiiciently large n, them is just one zero near each k, Since, according to (9.1), the functions y and z have the same positive zeros, it is sufficient to prove this for the hct ion

Page 228: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

z. If for arbitrarily large n, there wen a pair of zeros of the function z near the point k,,, then it would follow from Rolle's theorem that z' has a zero near z,,. But by (9.18) this is impossible, since near z,, = m - a, the value of z' is near A cos nn, provided that n is sufficiently large. Thus, for sufficiently large n, all the zeros of the function y lie near the numbers k , and there is only one zero near each k,,. It follows that the distance between consecutive zeros of the function y approaches x as the distance from the origin increases. In particular, these considerations apply to the functions Jp(x) and Yp(x), where according to (9.17), the numbers kn have the values

for Jp(x) and Yp(x), respectively. Below, we shall be concerned with the positive zeros of JJx). (It should

be noted that by (4.3), the positive and negative zeros of G(x) are located symmetrically with respect to the origin.) It follows from (7.2) and Rolle's theorem that there is at least one zero of the function x-pJHl(x) between any two consecutive zeros of the function x-pJ#, i.e., there is at least one zero of the function JP+l(x) between any two consecutive zeros of the function Jp(x). If we replace p by p + 1 in (7.1), we obtain

From this we conclude as before that there is at least one zero of the function Jp(x) between any two consecutive zeros of the function JHl(x). Thus, the zeros of Jp(x) and JH1(x) "separate each other." More precisely, one mul only one zero of JNl(x) appears between any two consecutive positive zeros of Jp(x). Moreover, Jp(x) and JHl(x) cannot have any positive zeros in common. In fact, if Jp(x) and Jp+l(x) both vanished at xo > 0, then by (7.8). J;(x) would also vanish at xo. But this is impossible, since by the uniqueness theorem for solutions of a second order differential equation, a x o ) = 0, J*) = 0 would imply that J'(x) r 0, which is obviously false.

Consider next the zeros of the function JL(x). By Rolle's theorem, at least one zero of Ji(x) lies between any two consecutive zeros of Jp(x). Therefore, like Jp(x), the function JL(x) has an infinite set of positive zeros.

Finally, we investigate the zeros of the function xlL(x) - HJ'), where H is a constant, a function which is often encountered in the applications. As we have just seen, the functions Jp(x) and Ji(x) cannot vanish simultane- ously (for x > 0). It follows at once that Jp(x) must change sign as we pass through a value of x for which Jp(x) vanishes. Let hl, A2, . . . , An, . . . denote

Page 229: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 10 BEssEL FUNCTIONS AND FOURIER-BESSEL SERIES 2 1 5

the positive zeros of J'(x) arranged in increasing order. For 0 < x < Al, the function Jp(x) does not change sign, and in fact for p > - 1 (the only values of p of interest to us), Jp(x) is positive for 0 < x < Al. [See formula (4.3), where the first term, which determines the sign of Jp(x) for x near zero, is positive.] As we pass through A,, Jp(x) goes from positive to negative values, as we pass through A2, Jp(x) goes from negative to positive values, etc., so that

Clearly, we have

Therefore, the function xJ@) - HJ,(x) is alternately negative and positive for x = Al, A2, S, . . . , and hence vanishes at least once between each pair of consecutive zeros AI, A,, A,, . . . . This shows that xl'(x) - HJp(x) also has an infinite set of positive zeros.

It can be shown that the distance between consecutive zeros of both JL(x) and xJ;(x) - HJJx) approaches x as the distance from the origin increases, just as in the case of the zeros of Jp(x). However, we shall not prove this fact here.

I I. Parametric Form of Bessel's Equation

Let the function y(x) be any solution of Bessel's equation (1.1). Consider the function y = y(Xx), and set 7t.x = t. Since obviously

substituting

into (1 1. I), and using hx = t, we obtain

Thus, if the function y(x) is a solution of Bessel's equation (1.1), the function y(k) is a solution of the equation

called the parametric form of Bessel's equation, with parameter A.

Page 230: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

12. Orthogonality of the Functions JP(Ax)

Let A and p be two nonnegative numbers. Consider the functions y = JP&) and z = Jp(p.x), where p > -1. According to Sec. 11, these hctions obey the equations

~ 2 3 ) ' + xy' + (Px2 - p2)y = 0,

x2z" + xz' + (~2x2 - pz), = 0

Subtract the first equation multiplied by z from the second equation multiplied by y. The result is

x(yzw - zy? + (yz' - Z Y ~ = ( ~ 2 - ~ ~ X Y Z ,

or

x(yd - zy')' + (yi - zy') = (A2 - p2)xyz, SO that finally

1M.V' - d' = G2 - p9xyz. (1 2.1)

We now intepate (12.1) from 0 to 1, obtaining

Since by hypothesis p > - 1, the integrand in (12.2) is actually integrable on the interval [O, 11. In fact, since y = Jm), z = Jp(p.x), it follows Rom (4.3) that

Y = -(XI, = xJW)s (1 2.3) where p(x) and Mx) are the sums of power series, and hence represent continuous finctions with continuous derivatives. Therefore

lxyzl = Ix-1 q(x)+(x)l < Mxwl (M = const),

and since 2p + 1 > -1 by hypothesis, this implies the integrability of the right-hand side (and hence of the lefi-hand side) of (12.1). It follows from (12.3) that

[XDd - zy?lp.o ' 0 for p > - 1. Therefore, instead of (12.2), we can write

Page 231: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 12 BESEL FUNCTIONS AND FOURIER-BESSEL SERIES 2 1 7

We now note that

h l = Jp(A). [zl*1 = Jp(p), while on the other hand

so that

1 = 9 [Yl, 1 = l*J;(p).

Thus, (12.4) becomes

So far, A and p have been arbitrary nonnegative numbers. We now impose certain restrictions on A and p. Consider the following three cases:

1) A and p are different zeros of the function Jp(x), i.e., Jp@) = 0, Jp(p) = 0, A # p. For such values of h and p, the left-hand side of (12.5) vanishes. Noting that - p2 # 0, we then obtain

If the factor x were absent in the integrand, the functions J P m ) and JP(px) would be orthogonal in the usual sense. In the present case, we say that the functions Jp(k) and Jp(px) are orthogonal with weight x. Of course, we might also say that the functions

are orthogonal in the usual sense. It should be noted that according to (8. l), for p = 1 12, zl and 22 reduce to

- zl = 421% sin X*, z2 = sin px,

where A and p are numbers of the form n According to (8.2), for p = - 112, we obtain the functions

where in this case A and p are numbers of the form (n + f)n. Thus, in these special cases, we obtain the trigonometric functions, i.e., in the first case we obtain the functions sin m x (except for a factor), which are orthogonal on [O,l], and in the second case the functions cos (n + f)m, which are also orthogonal on [0, 11.

Page 232: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

2) A and p are different zeros of the function Ji(x):

In this case, the left-hand side of (12.5) also vanishes, and hena we still obtain the formula (12.6). Thus, here again the functions J p ~ ) and JP(p) are orthogonal with weight r

3) Finally, let X and p be two different zaos of the function xJ;(x) - HJp(x) :

Multiply the first equation by Jp(p) and subtract it from the second equation multiplied by JP@). The result is

IJI,@Y'(p) - Mp(iJ.Y;(n) = o* Therefore, in this case, the left-hand side, of (12.5) also vanishes, and we again obtain the formula (12.6), i.e., the functions J Jk ) and Jp(px) are orthogonal with weight x.

13. Evaluation of the Integral 1' xJ;(~x) dx 0

When h and p are different, (12.5) leads to

When p + A, the fraction on the right becomes indeterminate, since the numerator and denominator both approach zero. To '' remove " this indeterminacy, we use L'Hospital's rule, letting A = const and p+ A. This gives

But Jp@), regarded as a function of A, satisfies Bessel's equation, i.e.,

Page 233: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 13 BESSEL FUNCTIONS AND FOURIER-BESSEL SERIES 2 19

so that

Therefore, it follows from (13.1) that

Thus, we can draw the following conclusions: 1) If A is a zero of the function Jp(A), then

This formula can be given a different form. We first replace x by A in (7.8), obtaining

Since in the present case Jp(A) = 0, we have

J;6 = - JHl(A), SO that

2) If A is a zero of the function J;(A), then

* 14. Bounds for the Integral I1 xJi(~x) dx 0

The following inequality, valid for sufficiently large A, will be useful later

Here K > 0 and M > 0 are constants (which can depend on p). Obviously we have

1: xJi(7a) dx = jt tJi(t) dt. (14.2)

According to the asymptotic formula (9.16)

Page 234: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

220 FUNCTIONS AND FOURIBR-BESSRL SERIES CHAP. 8

for sufficiently large t, and hence

In view of (14.2). this implies the right-hand inequality in (14.1). On the other hand, by the same asymptotic formula (9.16)

~ ( t ) = (A sin (t + 4 + i)' = A2 sin2 (t + o) + U r sin (t + o) r2

t + - t2

for large t, i.e., for t > )b. But then

and in view of (14.2), this implies the left-hand side of (14.1).

IS. Definition of Fourier-Bessel Series

The rest of this chapter is devoted to the study of Fourier expansions with respect to Bessel functions. Let hl, X2,. . ., )k, . . . be the positive roots of the equation Jp(x) = 0 ( p > - 1) arranged in increasing order. According to Sec. 12, the functions

form an orthogonal system on [0, I], with weight x. To give the reader some idea of the appearance of the functions of the system (15.1), in Fig. 46 we

Page 235: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Sac. 1s BEssEL PUNmONS AND FOUIUER-Bm SElUE3 22 1

draw the graphs of the functions Jl(Alx), Jl(A2x), JIG*) on the interval [O, 11. The functions Jl(A,p) (n = 3,4, . . .) on [O, 11 become more and more complicated, i.e., the number of "oscillations" increases.

For any function f (x) which is absolutely integrable on [0, 11, we can form the Fourier series with respect to the system (15.1), or briefly, the Fourier- Bessel series

where the constants

are called the Fourier-Bessel coeflcients of f(x). These coefficients can be obtained by using the following formal argument : Instead of (1 5.2), we write

We multiply both sides of (15.4) by xi l ' , ,~) and integrate over the interval [O, 11, assuming that term by term integration is justified. Because of the orthogonality (with weight x) of the system (15.1), the result is

which implies (1 5.3). [See equation (1 3.4).] If the equality (1 5.4) actually holds and if the series on the right converges

uniformly, then term by term integration is known to be legitimate, and hence the coefficients Cn must be given by the formula (1 5.3). However, just as in the case of ordinary orthogonal systems (see Ch. 2, Sec. 2), we first use (15.3) to form the series (15.2). and only later examine the Convergence of the series to f(x).

Id Criteria for the Convergence of Fourier-Bessel Series

We now state without proof the most important criteria for the con- vergence of a Fourier-Bessel series to the function from which it is formed. Thm criteria are analogous to those with which we are familiar in the case of trigonometric Fourier series (see Ch. 3, Secs. 9 and 12). However, the proofs are much more complicated than in the case of Fourier-Bessel series, and hence we omit them.

THEOREM 1. Let f(x) be a piecewise smooth, continuous or discon- tinuous function on [O, 11. Then the Fourier-Bessel series (p 3 - +) of

Page 236: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

222 BESSEL FUNCTIONS AND FOURIER-BESSEL SERIES CHAP. 8

f(x) converges for 0 < x < 1. Moreover, its sum equals f (x) at every point of continuity o f f (x) and f w(x + 0) + f (x - O)] at every point of discontinuity o f f (0).

The series always converges to zero for x = 1, and converges to zero for x = 0 if p > 0 (since all the functions of the system vanish for these values). We note that for p < 0 all the functions of the system (15.1) become infinite for x = 0 [see (4.1)], so that it is meaningless to talk of convergence of the series at x = 0.

Remark. Instead of piecewise smoothness of f(x) on [0, 11, it is sufficient to require piecewise smoothness of f (x) on every subinterval [a, 1 - 61 where 6 > 0, in addition to thewrequirement - that f(x) [or even &f(x)] be absolutely integrable on the whole interval [0, 11.

THEOREM 2. Let f (x) be continuous and have an absolutely integrable derivative on the interval [a, b], where 0 ( a < b ( 1. Then, the Fowier-Besel series (p 3 -+) of f (x) converges miformly on every subinterval [a + 6, b - 61, where 6 > 0.

THEOR~M 3. Let f(x) be absolutely integrable on [O, 11, let f(x) be continuous and have an absolutely integrable derivative on the interval [a, 11, where 0 g a < 1, and let f (x) satlsfV the condition f(1) = 0. Then the Fourier-Bessel series ( p 2 - f) o f f (x) converges un$iormly on every subinterval [a + 6, 11, where 6 > 0.

The condition f(1) = 0 is quite natural, since all the functions of the system ( 1 5.1) vanish for x = 1.

Remark. In Theorems 2 and 3, it is sufficient to require only absolute integrability of z/;f(x).

Example. Expand the fiction f(x) = x p (p 2 - f) for 0 < x < 1 in Fourier-Bessel series with respect to the system

By formula ( 1 5.3)

But we have

According to formula (7.1) (with p replaced by p + 1 )

Page 237: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Therefore

so that 1 J: x p + l J p ( ~ ) dX = - JIG (A,,).

An

It follows that

Thus, by Theorem 1, we can write

* 17. Bessel's Inequality and Its Consequences

The orthogonality with weight x of the functions of the system (15.1) can be regarded as ordinary orthogonality of the functions

Therefore, if we wish to make a series expansion of a function f(x) with respect to the system (15.1), we can fitst make a series expansion of the function fif(x) with respect to the ordinary orthogonal system (17.1), obtaining

and then go over to the expansion (15.2). (It is easily verified that the coefficients of both expansions are the same.) These considerations allow us to apply the results of Ch. 2 to Fourier-Bessel series.

Thus, assuming that F(x) = dif (x) is a square integrable function (which is certainly the case if f(x) itself is square integrable) and applying Bessel's inequality (see Ch. 2, Sec. 6), we obtain

Page 238: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Consequently, we have

lim [c: I1 XI;@&) &] = 0. H a 0 0

But according to (14.1)

for all sufficiently large n, and therefore 2

cn lim - = 0 -00 A,

c, lim - = 0. -a

Even more can be said: It follows from the asymptotic formula (9.16) that

for all sufficiently large x, and hence for every fixed x > 0

if n is sufficiently large. Then, using (17.2). we obtain

for every Bxed x > 0. Thus, if f(x) is square integrable, the general tenn of the Fourier-Beweel series of f(x) always converges to zero (x > 0). If p > 0, then obviously this will also be the case for x = 0 (since all the functions of the system (15.1) vanish for x = 0). Finally, it follows from (15.3) that

for sufficiently large n, so that by (1 7.2)

Page 239: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

mi& i s the analog of the property of the trigonometric integrals discussed in Ch. 3, Sec. 2.

Equations (17.2) and (17.5) are true not only for square integrable functions, but also for arbitrary absolutely integrable functions. (We omit the proof.)

* 18. The Order of Magnitude of the Coefficients which Guarantees Uniform Convergence of a Fourier-Bessel Series

THEOREM 1. Ifp 3 0 and

where G > 0 and c are constants, then the series

converges absoIuteIy and unifomly on [0, 11.

Proof: Forp 3 0, the function Jp(x) is bounded in the neighborhood of x = 0. By the asymptotic formula (9.16), Jp(x) is bounded for large r Therefore, Jp(x) is bounded for aN (positive) x, i.e.,

1 ( A ) n (L a const),

so that by (1 8.1)

But since - A,,+n as n + co (see Sec. 9), it follows that for n > m (where m is some fixed number),

Therefore, if n is large, A,, 3 +n and

Consequently, for large n we have

H IcnJp(ht41 c ;;i~; (H = const).

Theorem 1 now follows from the fact that the right-hand side of this inequality is the general term of a convergent numerical series.

Page 240: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

for all mficiently large n, where E > 0 cmd c are constants, then the series

converges absolutely and w@wmly on the interval [O, 11. nis implies that the series (4.2) converges absolutely and u e r m l y on every sub- interval [8, 11, where 8 > 0.

Proof. For p 3 -& the function I& J,,(x) is bounded as x+ 0 [see equation (4.3) 1. According to the asymptotic formula (9.16).

JP(x) is bounded for large r Therefore, Jp(x) is bounded for all (positive) x, i.e.,

and hence for any x in [O,1] we have

It follows that

H I c n d i ~ p @ & l * .1+= (H = const).

The right-hand side of (18.7) is the general term of a convergent numer- ical series. This proves the first part of the theorem. The second part of the theorem is a consequence of the following inequality, obtained from (18.7) :

for all mficiently large n, where E > 0 and c are constants, then the series (18.2) converges absolutely and mf fon ly on every interval [8,1], where 8 > 0.

Page 241: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 18 BESSEL FUNCTIONS AND FOURIER-BESSEL SERIES 227

Proof. Let 6 6 x ( 1. By the asymptotic formula (9.16),

for all sufiiciently large x, i.e., for x 2 xo. Moreover, the inequality An6 3 xo holds for all sufficiently large n. Therefore if x 3 8, we have l,p a xo a fortiori, and hence for 6 ( x ( 1

so that

or

By (1 8.3), this implies

for all sufficiently large x. Theorem 3 now follows from the fact that the right-hand side of this series is the general term of a convergent numerical series.

Remark. In the conditions (18.1) and (18.4) of Theorems 1,2, and 3, we can write simply n instead of AR [in view of (18.311.

Example 1. The series

is absolutely and uniformly convergent on [O, 11, since here p = 1, E = 1, and hence Theorem 1 is applicable.

Example 2. The series

is absolutely and uniformly convergent on every interval [8, 11, 6 > 0, while the series

Page 242: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

is absolutely and uniformly convergent on the whole interval [O, 11. Here Theorem 2 is applicable.

Example 3. The series

is absolutely and uniformly convergent on every interval [a, 11, 8 > 0. Here Theorem 3 is applicable.

*19. The Order of Magnitude of the Fourier-Bessel Coefficients of a Twice Differentiable Function

LWA. Let F(x) be a twice diflerentiable fmction defined on the interval [O, 11, such that F(0) = F'(0) = 0, F(1) = 0, and such that F"(x) 25 bounded (the second derivative may not exist at certah points). men, i f X ir a zero of the f i o n Jp(x), where p > - 1 , the inequality

IJ: fi F ( ~ ) J ~ ( w & (R = const)

holds.

Proof. According to equations (9.1) and (9.2). the function

z(t) = fi ~'(t) satisfies the equation

If we set t = Xw, then

so that

or

Page 243: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Thus, the function r = ds ~~(7a) satisfies the equation (19.2). But then the function z = 6 x J&) also satisfies (19.2), since it differs from fi J p ~ ) only by a constant factor.

We now set p2 - = m in (1 9.2) obtaining

It follows that

Since

(F'z - Fz')' = F.r - FzW,

we have

1 1 = 12 - J 0 ( ~ ( x ) - P(x)) z tix + [F'z - FZ'J;:;.

But

[F'z - FZ']::: = [F8(l)z(l) - F(l)z8(l) ] - [Ff(0)z(O) - F(O)f(O) ] = 0,

in view of the following facts:

1) z(1) = AX)^-^ = Jp@) = 0 and F'(1) is finite; 2) F(l) = 0 by hypothesis and zt(l) is finite; 3) By Taylor's formula, Ff(x) = xFU(Ox) (0 < 0 < l) , while

z = f i J p ( k ) = xp+clf2)9(x), where p(x) is continuous and differenti- able, being the sum of a power series [see equation (4.3)J. Therefore

Ff(0)z(O) = lim F8(x)z(x) = lirn x*(3~2)p(x)FW(0x) = 0 X--+O -0

since F" is bounded and p > - 1 ;

4) By Taylor's formula, F(x) = )x2F"(Ox) for 0 < 8 < 1, while

zt (x) = (x*(lf2)p(x))'

= 0, + +)xp(ll2)p(x) + xr+'"2$'(x), SO that

F(O)z'(O) = lirn F(x)zr(x) A

= -) lim [(p + ))XP+(~/%(X) + xp+(s/2)9'(x)JF"@x) = 0. x 4

Page 244: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Using all these facts, we find that

As we have already noted, it follows from Taylor's formula that

near x = 0. This implies that the integrand of (19.3) is bounded. But then

Thus, by the Schwan inequality [see equation (4.1) of Ch. 21

M (J: lzl dx)f G J1 o z2 dk = [:d;(~#) dx 4 (M = const)

[see also equation (14.1)], and therefore

If we use (19.4). it follows from (19.3) that

which proves the inhquality (19.1).

THEOREM 1. Let f(x) be a b o d e d and twice dzrerentiable function defied on the interval [O, 11, such that f(0) = f'(0) = 0, f(1) = 0, and such thatf"(x) is bounded (the second derivathe may not exist at certain points). Then the Fwier-BesseZ coeficients of the f ~ i o n f(x) sathfy the inequality I

C Ic.1 4 (C = const).

Proof. If f(x) satisfies the conditions of the theorem, so does the function F(x) = l/ji f (x). Therefore, applying the lemma, we obtain

R d- (R = const). q2

By equation (1 4.1) 1 I, xJ;(h&) L An (K g O),

Page 245: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 19

so that

which proves the required inequality (19.5).

Remark. Theorem 1 remains true if we impose the requirements of the theorem on the function F(x) = fi f(x) instead of on f(x), since we actually applied the lemma to F(x).

The following proposition supplementing Theorem 2 of Sec. 16 is a consequence of Theorem 1 :

THEOREM 2. Let f(x) be a function which is continuous and twice differenthble on the interval (0, I ] , let f (0) = f'(0) = 0, f (1) = 0, and letr(x) be bounded (the second derivative may not exist at certain points). Then, the Fourier-Bessel series o f f (x) converges absoluteIy and la firmly on every subinterval 18, 11 where 0 < 6 if p > - 1, and on the whole interval [0, 11 I f p 3 0.

Proof. If p > - 1 , the assertion follows from the preceding theorem and Theorem 3 of Sec. 18. If p 2 0, the uniform convergence on the whole interval [0, 11 follows from the preceding theorem and Theorem 1 of Sec. 18.

Remark. I f we use Theorem 2 of Sec. 18, then with these conditions on f(x) and with p 2 - +, we obtain absolute and uniform convergence of the series (18.5) on the whole interval [O, 11.

*20. The Order of Magnitude of the Fourier-Bessel Coefficients of a Function Which is Differentiable Several Times

THEOREM 1. Let f(x) be a fmction defied on the interval [ O , l ] such that f(x) is dz&j7erentiable 2s times (s > 1) and such that

1 ) f(0) =f ' (O) = = fcb-l)(O) = 0 ; 2) ftb)(xJ is bounded (this derivative may not exm at certain points); 3) f(1) = f(1) = * * * - f(b-Z)(l) = 0.

Then the following hequality is satkfied by the Fourier-Bessel coeflcients o f f (0) :

C Icnl C ~ Z r ( I , S ) (C = const). (20.1)

Page 246: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Proof. It is easy to see that the function F(x) = fi f(x) also satisfies the conditions of the theorem. In particular, F(x) satisfies the conditions of the lemma of Sec. 19 and hence satisfies (19.3). im.,

where m = 3 - 3, z = &J~A,,X). If Fl denotes the function in parentheses in the last integral, then we have

Since the function Fl satisfies all the conditions of the lemma, this time (19.3) gives

where we have written

If s > 2, then F2 again satisfies the conditions of the lemma, and in fact we can repeat the argument s times, finally obtaining

where

is a bounded function. It follows that

so that we have

Page 247: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 20

But

and since

according to equation (14.1) we finally obtain

which proves the inequality (20.1).

The next result is a consequence of Theorem 1 :

THEOREM 2. If the hypotheses of Theorem 1 are met for s 3 1, then 1) For p 3 0

H I CJP(~~X) I ( ),f:-U,z) (H = const)

for any x (0 < x < 1);

2) Forp 2 -)

L IcnJp(Xm4I 6 .\Tx (L = const)

for all x (0 < x ( 1) unlfrmly;

3) For p > - 1, the inequality (20.3) holdc for each x (0 < x < 1) i f n > n(x). (In this case, there is no unifrmity in x.)

Proof. In Sec. 18 (see the proof of Theorem l), we saw that the function Jp(x) is bounded for p 3 0. Hence, the inequality (20.2) is an immediate consequence of (20.1). For p 2 - +, since Jp(Anx) satisfies the inequality (18.6), we need only apply (20.1) to get (20.3). Finally, for p > - 1, it follows from the asymptotic formula (9.16) that

I

'-L IJpob9I 6 7 (L = const),

Anx

for every x (0 < x $ 1) and n > n(x). Again using (20.1), we obtain (20.3).

Page 248: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

*21. Term by Term Differentiation of Fourier-Bessel Series

Given a Fourier-Bessel series

we now find sufl6cient conditions for the validity bf the equality

It follows from formula (7.8) that

We assume that p > -- 1, so that p + 1 > 0. Therefore, the quantity

is bounded, by the asymptotic formula (9.16), and hen=

where we consider only values of x such that 0 d x d 1. If

where s > 0 and C are constants, then for x > 0

This implies at once [see (18.3)] that the series (21.2) converges for 0 < x 4 1 and converges uniformly on every subinterval [a, 11 where 0 < 8 < 1. The last fact implies the validity of (21.2) for 0 < x d 1.

As for the validity of (21.2) for x = 0, if p < 1, p # 0, it is easy to see that all the functions J;(A,,x) become infinite for x = 0, since Jp(x) = xRcp(x), where cp(x) is differentiable and cp(0) # 0 [see (4.311, and therefore in this case (1 8.2) becomes meaningless. If p 2 1, then (7.9) implies that

Page 249: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 21 BESSEL FUNCTIONS AND FOURIER-BESSEL SERIES 235

where the functions in the right-hand side have nonnegative indices and hence are bounded. Then

I cJJi(As)l 6 I c A I H (H = const),

so that if

the series (21.2) is uniformly convergent [cf. (18.3)) on [O, 11, i.e., (21.2) holds everywhere on the interval [O, 11. Finally, if p = 0 we have

instead of (21.3). Since the function JI is bounded, then

if (21.6) holds, and the series in (21.2) is again uniformly convergent, so that (21.2) is valid for all x in the interval [0, 11.

Thus, finally, we have proved the following theorem:

THEOREM 1. IfP z -- 1 and Vc. sattsfies the condition (21 .S), then the s e r k (21.1) can be differentiated term by term for 0 < x < 1. If p = 0 or p a 1 and Cn satwes the condition (21.6). then the series (21.1) can be differentiated term by term everywhere on [O,l].3 We now find a sufficient condition for differentiating the series (21.1)

twice term by term, i.e., for the validity of the relation

Since Jp(x) is a solution of Bessel's equation, we have

h>2q@s) + uJ;w) + #x2 - pWp(h,P) = 0-

Thus

so that by (21.4)

3 Note that the convergence of the series (21.1) itself follows from (21.5) and (21.6) and the theorems of Sec. 18.

Page 250: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Therefore, we have

where c > 0 and C are constants, then by (20a4), we obtain

This together with (18.3) implies the convergence of the series in (21.7) on every interval [8,1], where 0 < 8 < 1. Since the condition (21.8) implies the convergence of the series (21.2) for 0 < x ( 1, the uniform convergence of the series in (21.7) on every interval [a, 11 implies that (21.7) is valid for O < x g l .

Next we observe that for - I < p < 2, p # 0, p # 1, all the ftnctions .$(A,,#) become infinite at x = 0, since .Ip(#) = x ~ ( x ) , where p(x) is the sum of a power series [p(O) # O] and hence is differentiable any number of times [see equation (4.3)]. Therefore, in this case the relation (21.7) becomes meaningless. Finally, we note that an argument similar to that given in connection with Theorem 1 can be used to show that for p 3 2, p = 0 or p = 1 and for

where c > 0 and C are constants), the relation (21.7) holds everywhere on [0,1]. Thus we have

THEOREM 2. u p > - 1 and if the coeflcients c, sathfy the inequality (21.8), then the series (21 .l) can be diflerentttated twice term by term for 0 < x ( 1.4 Ifp = 0,p = 1 orp 3 2 ,d i f thec , sa t i $ f y thecondition (21.9), then term by term dffJerentiation is possible everywhere on [0, 11. As a corollary of Theorem 2, we obtain

THEOREM 3. If the hypotheses of Theorem 1 of Sec. 20 are met for s = 2, then the Fourier-Besseel series of the function f(x) cun be difjren- tiated twice term by term jbr 0 < x < 1 i f p > - 1 and everywhere on [O, 11 i f p = 0, p = 1 or p 2 2.

4 The convergence of the series (21 .I) itself follows from (21.8) or (21.9) and the theorems of Sec. 18.

Page 251: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

sac. 22 BES~EL FUNCTIONS AND FOURIER-BESSEL SERIES 237

In fact, in this case, according to (20.1)

C Ic.1 7 ~ ~ 1 2 (C = const),

and we need only apply Theorem 2.

22. Fourier-Bessel Series of the Second Type

We now let Al, Az, . . . , A,,, . . . be the roots of the equation

arranged in increasing order. For H = 0, this equation becomes

The existence of an infinite set of positive roots of equation (22.1) [and in particular of equation (22.2)] was proved in Sec. 10. Moreover, according to Sec. 12, for p > - 1 the functions

form an orthogonal system on [0, 11, with weight x. The considerations given below pertain to the case where the A, are the roots of (22.1); all the results are valid for the special case where the system (22.3) corresponds to the roots of (22.2).

Let f(x) be any function which is absolutely integrable on [0, 11. Then we can form the Fourier series of f(x) with respect to the system (22.3). We shall call such a series

a Fourier-Bessel series of the second type; here the constants

just like ordinary Fourier-Bessel coefficients (or for that matter, just like Fourier coefficients in general), can be obtained by the usual formal argument. We have the following theorem, which closely resembles Theorem 1 of Sec. 16 :

THEOREM. Let f(x) be a piecewise smooth (continuous or discon- tinuous) function de9ned on [O, 11. l k n the Fourier-Bessel series of

Page 252: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

f (x) of the second type ( p 3 - +, p > H) converges for 0 < x < 1. More- over, its sum equals f (x) at every point of continuity of f(x) and

at every point of dbcontinuity off (x).

For x = 1, the series converges to the value f(1 - 0); if f(x) is con- tinuous at x = 1, the series converges to f (1). If p > 0, the series con- verges to zero, since in this case all the functions of the system (22.3) vanish at zero. Since all the functions (22.3) become infinite at x = 0, it is meaning- less to talk about the convergence of the series at x = 0.

We omit the proof of this theorem, noting only that the condition p > H (which has no analog in the theorems of Sec. 18) is required because of complications that arise when p < H. It turns out that if p 6 H, then the theorem is true only if we add a new function to the system (22.3). For example, if p = H the new function is xp, and instead of (22.3) we have to consider the new orthogonal system

In this case, the orthogonality of the function x p (with weight 1) to the other functions is a consequence of the following two facts:

1) x = 0 is a root of the equation

xJ'(x) - pJp(x) = 0

[see equation (7.8)]; ' 2) xp satisfies Bessel's equation with parameter A = 0, i.e., the equation

x2yW + xy' - p2y = 0,

as can be verified directly. Therefore, we can apply to the functions xp and JP(h~) all the considera-

tions of Sec. 12 (where the orthogonality of the solutions of the parametric form of Bessel's equation was discussed). For p < H, the "additional" function is much more complicated. Thus, for simplicity, we restrict ourselves to the case p > H. A remark like that made after the statement of Theorem 1 of Sec. 16 also applies to the present theorem, and moreover, there are also two theorems completely analogous to Theorems 2 and 3 of Sec. 16, with the supplementary condition p > H in both cases and the elimination of the condition f (1) = 0 in the second case. (The remark made in connection with Theorems 2 and 3 of Sec. 16 remains in force.)

Example. Make a series expansion of the function f (x) = xp (0 < x ( 1) with respect to the system

JpoW), Jp(A2~1, 9 JP(h&), s (22.6)

where the )k are the roots of the equation J;(x) = 0 (p > 0).

Page 253: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

By formula (22.5)

cn = 2" 1' xp+1Jp()CT) dx, - P * 0

and by (16.1)

Therefore, by the theorem given above, we can write

We now ask whether this expansion holds for p = 0. The answer is negative, since for p = 0

[where we have used equation (7.8)]. Thus, the right-hand side of (22.7) is zero, while the left-hand side is f(x) = xO = 1. The reason why the expansion fails for p = 0 is the following: In our case H = 0, so that p = H. Then, as remarked above, the system (22.6) must be supplemented by the function xP = XO = 1. If we denote the coefficient corresponding to this function by co, we obtain

since f(x) = 1. Moreover, Cn = 0 if n = 1,2,. . .. Therefore, instead of (22.7) we obtain the expansion

which is obviously valid!

"23. Extension of the Results of Secs, 17-21 to Fourier-Bessel Series of the Second Type

In the theorems of Secs. 17 and 18, we essentially made no use whatsoever of the assumption that the numbers )k are the roots of the equation Jp(x) = 0. Therefore, these theorems are equally true for Fourier-Bessel series of the second type. However, in the lemma of Sec. 19, we did use the condition JP() = 0 to prove the formula (19.3). Thus, we now need a new lemma, which we proceed to prove.

Page 254: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

240 BBssgL FUNCTIONS AND FOURIER-BES8BL SBRIES CHAP. 8

LEMMA* Let F(x) be a twice dzfereetiable function de$ned on [0,1], w h that F(0) = Ft(0) = 0, Ft(l) - (H + +)F(l) = 0, and such that Fu(x) Is b o d e d (the second derivative may not exfit at certain points). men, if A is a root of the equation

where p > - 1, it follows that

R I = l ~ ~ f i ~ ( x ~ ~ ) ~ l G (R = const). (23.1)

Proof. We arrive at the equality

just as in the lemma of Sec. 19, and the whole problem reduces to showing that the last term vanishes. This last term is the difference

We begin by calculating the first term in brackets. Since z = ~&~@x) , we have

where we have used the condition N'@) - HJpQ = 0. But then the first term in brackets is just

which vanishes by the hypothesis of the lemma. The fact that the second term in brackets in (23.2) vanishes, and indeed the rest of the proof follows by just the same method as in the lemma of Sec. 19.

The lemma just proved leads to the following result:

THEOREM 1. Let /(x) be a twice differentiable function defied on the t n t d [0, 11, such that f(0) = f (0) = 0, f (1) - Hf(1) = 0,s and such

The condition f'(1) - Hf(1) = 0 appears artificial at this point, but it arises quite naturally in the applications.

Page 255: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

that f (x) is bounded (the second derivative may not exist at certain points). Then the inequality

C IcnI 9 (C = const), (23.3)

n

is obeyed by the Fourier coeficients of f(x) with respect to the system (22.3).

Proof. If F(x) = <x f(x), then obviously F(0) = F(0) = 0 and Fff(x) is bounded. Moreover

f (x) F'(x) = + fi f (x),

so that

F(1) - (H + +)F(l) = +j'(l) +f(l) - (H + +)f(l) -f'(l) - Hf(1) = 0.

Thus, the lemma can be applied to the function F(x); the result is

Since by formula (14.1)

we obtain (23.3), taking account of (22.5).

Theorem 1 and the results of Sec. 18 imply

TReo~ear 2 Let f(x) be a twice &firentiable function on the interval [0,1], such that f(0) = f (0) = 0, f(1) - Hf(1) = 0 and such that f "(x) is boundkd (the second derivative may not exist at certain points). Then the Fourier-Bessel series of f(x) of the second type converges abso- lutely and unifrmly on every subintervaZ [a, 11, where 0 < 8 < 1, rp fp - 1 and on the whole intevval [0, 1) vp fp 0.6

Finally Theorems 1 and 2 of Sec. 21 apply equally well to Fourier-Bessel series of the s+oond type.

24. Fourier-Bessel Expansions of Functions Defined on the Interval [0, I ]

Let f(x) be an absolutely integrable function defined on the interval [O, I ] , and set x = it or t = x/l. Then the function p(t) = f(lt) is defined on

6 If the condition p > H appearing in the theorem of Sec. 22 is not met, then the serie need not have f(x) as its sum.

Page 256: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

the interval [0,1] of the tgxis, and we can write

where

in the case of a Fourier-Bessel series of the first type, or

in the case of a Fourier-Bessel series of the second type. Returning to the variable x, we obtain

f (x ) - C ~ J , ( ~ X ) + cdp(?x) +*-+ eJp ("x) + * b e , (24.2)

where

If the series (24.1) converges, then the series (24.2) converges, and conversely. It should be noted that we can also obtain the expansion (24.2) directly,

avoiding the auxiliary function cp(t), if we observe that the system

is orthogonal with weight x on the interval [0, 11. In fact, we have

where we have changed the variable of integration by setting x = It. Once we have established the orthogonality of the system (24.5), we calculate the Fourier coefficients in the usual way, obtaining (24.3) or (24.4). depending on whether the numbers A,, are the roots of the equation Jp(x) = 0 or of the equation xl'(x) - HJp(x) = 0. Just as in the case of the interval [0, 11, the &es (24.2) with the coefficients (24.3) is called a Fourier-Bessel series of the $rst type, and the wries (24.2) with the coefficients (24.4) is called a Fourier- Bessel series of the second type. Since we can go from the series (24.2) to the

Page 257: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

series (24.1) by making the substitution x = Zt, all our previous results for the interval [0, 11 also hold for the case of the interval [0, I]. In particular, the convergence criteria ars valid, provided of course that they are formulated for the interval [0, I ] instead of for [O, 11.

PROBLEMS

1. Write the general solution of the differential equation

5 y" + ,y8+ y = 0.

2. Calculate

where n is a positive integer.

3. Find an asymptotic formula for Jb(x). Hint. Use fonnulas (7.9) and (9.16).

4. Find an asymptotic formula for J;Xx). Hint. Use equation (I .I), fomula (9.16) and the preceding problem.

5. Show that the hctions Jp(x) (p 3 0) and fi JJX) (p 5 - f) are bounded f o r O < x < 00.

Hint. Use formulas (4.3) and (9.16).

6. Show that

7. By multiplying the power series for eM/2 and e-xia, show that

8. By substituting t = c" in the formula of the preceding problem, taking real and imaginary parts, and using the formula for trigonometric Fourier ooeffllcients, show that

(n = 0, 1,2,. . .) J*+I(X) = r sin (x sin 0) dn d0.

It 0

9. Show that

a) I ~ J ~ ( X ) I C ~ forallx (k,n=O,I,* ,... );

Page 258: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

c) lim Jn(x) = 0 for all X. ))--*OD

10. Expand the function x-p (0 < x < 1) in Fourier-Wl series of the first kind.

11, Expand the function xp (0 < x < 1) in Fourier-BessA series with respect to the system

where the A, are the roots of the equation xJL(x) - HJp(x) = 0. Hint. Use the example in Sec. 22.

12. Expand the function x3 (0 6 x < 2) in Fourier-Bessel series of the first kind, with p = 3.

Hht. Use formula (24.3) and the results of the example in Sec. 16.

13, Let Al, A,, . . . be the positive roots of the equation Jo(x) = 0. Show that

Hint. Use formulas (7.1), (7.2)' and (15.3).

14. Let Ax,, A2,. . . be the positive roots of the equation Jp(x) = 0 ( p > - f). Show that

(Ineachcase,O 6 x 6 1.) Hint. a) Multiply the last formula of Sec. 16 by xp+l, and integrate from 0

to x, using (7.1); b) Multiply (I) by X P + ~ and integrate; c) Multiply (I) by r ( p + l ) and differentiate, using (7.2).

Page 259: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

THE EIGENFUNCTION METHOD AND ITS APPLICATIONS

TO MATHEMATICAL PHYSICS

Part I. THEORY

I. The Gist of the Method

Many problems of mathematical physics lead to linear partial differential equations. Examples of such equations are

where P, A, Q are continuous functions of the variable x, and u = u(x, t ) is an unknown function of the variables x and t. The first equation arises in problems involving vibrations of strings and rods, while the second arises in problems of heat flow, We shall confine our attention to equation (1.1). since this will be quite sufficient to explain the gist of the method to k discussed here.

In every concrete problem leading to an equation of the form (1.1), one requires a solution u of (I. 1) which satisfies certain conditions. For example, suppose that we have to find a solution u = u(x, t), defined for a 6 x < b and all t 2 0, which satisfies the boundary conditions

Page 260: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

246 THE BIGENFUNCTION METHOD AND ITS APPLICATIONS CHAP. 9

for any t 2 0, where a, p, y, and 6 are constants, and which satisfies the initial conck'tions

for a ( x $ b, where f (x) and g(x) are given continuous functions. Usually, x stands for length and t for time, which explains the terms boundary condi- tions and initial conditions. We shall assume that neither the pair a, nor the pair y, 6 can vanish simultaneously [for otherwise, instead of the relations (1.3), we would have the vacuous identities 0 - 01. This assumption can be written as

To solve the equation (1.1). we k t look for particular solutions of the form1

which satisfy only the boundary conditions (1.3). (We are only interested in solutions which do not vanish identically.) With this in mind, we differ- entiate (1.6) and substitute the result in (1.1); obtaining

whence

Since the left-hand side of this last equation is a function only of x, while the right-hand side is a function of t, the equality is possible only if both sides equal a constant :

This leads to the following two ordinary linear differential equations of the second order:

PCP' + A@' + Q@ = -A@, T"+hT=O.

Obviously, a function u @ 0 of the form (1.6) satisfies the boundary condi- tions (1.3) if and only if the function @(x) satisfies the boundary conditions

1 This is known as the method of seporaion of tmirrbles. (~unslotor)

Page 261: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

The problem of finding a solution of the equation (1.7) satisfying the condi- tions (1.9) will be called the bounhry value problem for the equation (1.7) with the conditions (1.9).

In general, the boundary value problem for a linear differential equation of the second order does not have a solution for every value of A; in particular, this is true of the boundary value problem for the equation (1.7) with the conditions (1.9). Nevertheless, it can be shown that there exists an inflnite set of values Ao, At, . . ., A , . . . for which the boundary value problem has a solution, provided only that P # 0. Every value A for which the boundary value problem has a solution 6 $ 0 is called an eigenvalue, and the solution 6 corresponding to A is called an eigenfmction. It will be shown below that in our case only one eigenfunction corresponds to each eigenvalue (to within a constant factor). Thus, for our problem, there is an infinite set of eigen- values Ao, Al, . . . , An, . . . , with corresponding eigenfunctions

As will be shown in Sec. 4, the functions (1.10) form an orthogonal system on [a, b], with a certain weight.

Having finished with equation (1.7), we next solve equation (1.8) for each A = A,, and find the corresponding function T,(t), which depends on two arbitrary constants A, and Bn. Thus, if )k > 0 for n = 0, 1,2, . . . (and this case occurs quite often in concrete problems), we obviously have

Tn(t) = An cos fin r + 4 sin z/G t, (1.11)

where An and Bn are arbitrary constants. Then, each function

will be a solution of equation (1.1) satisfying the boundary conditions (1.3). Because of the linearity and homogeneity (with respect to u and its derivatives) of equation (1.1), every finite sum of solutions of (1.1) is also a solution. The same is also true of the infinite series

if it converges and if it can be differentiated term by term twice with respect to x and t. If this is the case, we have

Page 262: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

248 THE EIOENFUNCTION METHOD AND ITS APPLICATIONS CHAP. 9

Since each term in parentheses in the last sum-vanishes because Un is a solution of (l.l)Jy the entire sum vanishes, which means that the function (1.12) is a solution of (1 .I). Moreover, since each term of the series (1.12) satisfies the boundary conditions (1.3), the sum of the series, i.e., the function u, also satisfies these conditions.

We must now satisfy the initial conditions (1.4). This can be achieved by suitably choosing the values of the constants An and Bn appearing in the expressions for the functions un(xY t). With this in mind, we require that the relations

hold, which is equivalent to requiring that the functions f(x) and g(x) can be expanded in series with respect to the eigenfunctions (1.10). The possi- bility of making such expansions can be proved under rather broad conditions on the coefficients in equation (1.1) and on the functions which are to be expanded. Thus let

Then we need only set

in order to find An and 4. Hence, if (1.1 1) holds, we have

so that

Our results are based on the supposition that the series (1.12) converges and can be differentiated term by tenn twice with respect to x and t. There- fore, the coefficients An and Bn just found must be such as to guarantee that

Page 263: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 1 THE EIGENFUNCTION METHOD AND rn APPLICATIONS 249

this is the case. However, in actual problems, the coefficients A, and B,, often do not have this property. Whether or not the series (1.12) will converge in such a case will be discussed in Sec. 7. In the meantime, we make the following remarks :

We know that series often define discontinuous functions. Hence, in order to avoid confusion, a few words concerning boundary conditions and initial conditions are in order. By the conditions (1.3) and (1.4), we mean more precisely

a lim u(x, t) + lim Wx, t) = 0, X* - ax

y lim u(x, t) + 6 lim Wx, t) = O x+b x+b ax

instead of (1.3), and

lim u(x, t ) = f(x), lim au(x, t) = b(x) t-0 t+o at

instead of (1.4). In other words, in (1.3) and (1.4), the values of #(a, t), au(a, t)/ax, etc., are to be interpreted as the limits to which u(x, t), h(x, t)/ax, etc., converge as the point (x, t) lying in the region a < x < b, t > 0 con- verges to the corresponding boundary point. It is quite clear that only such an interpretation of the boundary and initial conditions can correspond to the physical content of the problem. In the same way, when we say that the function u(x, t) is continuous in the region a < x ( b, t 2 0, we mean that u(x, t) is continuous in the region a < x < b, t > 0 and that the limit

lim U(X, t) (a < x < b, t > 0) X+X0 '-*to

has a finite value for every point (xo, to) on the boundary of the region.2 Then, it is easy to show that the boundary conditions vary continuously as the point (xo, to) is moved along the boundary.

Subsequently, by the solution of equation (1.1), or of some similar equa- tion, we shall always mean a solution which is continuous in the sense just indicated. It is easy to see that if such a solution is to exist, then the bound- ary conditions and the initial conditions must "agree" at the points (a, 0) and (b, 0) in such a way that the boundary values do not undergo a dis- continuity. Thus, for example, if the conditions (1.3) have the form u(0, t) = 0, u(1, t) = 0, and the conditions (1.4) have the form u(x, 0) = x + 1, au(x, O)/at = x2, then it is clear that the boundary values undergo a discontinuity at the points (0,O) and (1, 0), so that the problem certainly cannot have a continuous solution.

2 The analytic expression for u(x, t ) may not be a continuous function, i.e., may undergo a jump on the boundary.

Page 264: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

250 THE EIG~UNCTION ~ O D AND ITS APPLICA~ONS CHAP. 9

2. The Usual Statement of the Boundary Value Problem

We shall assume that the function P in equation (1.1) does not vankh. Obviously, equation (1.7) neither loses nor gains eigenvalues and eigen- functions if we multiply all its terms by a nonvanishing function. We now show that by carrying out such a multiplication, we can transform (1.7) into the form

o ' + q @ = --m, (2.1)

where p, q, and r are continuous functions of x on [a, b], p is a positive fun* tion with a continuous derivative, and r is a positive function.

Proof: First we assume that P > 0. (This can be done without loss of generality, shoe otherwise we need only multiply all the terms of (1.7) by - 1 and replace - A by x.) Then we solve the system

obtaining

where xo is some point of the interval [a, b]. (We take the constant of integration to be zero.) Obviously, p > 0, p' is continuous and r > 0. Now, we n e d only consider

rP@' + rR@' + rQ@ = -A&

and set

q = rQ. Then, amrding to (2.2),

+ p'@' + q@ = -ha, which is just the desired equation (2.1).

The boundary value problem is usually posed for an equation of the form (2. l), with coefficients satisfying the requirements just stated. The boundary conditions are still given by the same formulas (1.9).

3. The Existence of Eigenvalues

We shall not give a complete proof of the existence of eigenvalues for the boundary value problem under consideration; instead we just describe the basic idea of such a proof. Thus, in equation (2.1) we give A a bed value (real or complex) and find a solution satisfying the conditions

Page 265: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

We denote this solution by Qyx, A). Obviously

i.e., @(x, A) satisfies the first of the boundary conditions (1.9). (The prime denotes differentiation with respect to x.) As A changes, @(x, A) also changes, but nevertheless continues to satisfy the condition (3.1). Thus, a known function of x and the parameter A satisfies the &st of the conditions (1.9) for any A. (In the theory of differential equations, it is shown that @(x, A) can be represented in the form of a power series in A, and hence is an analytic function of A for all values of A.)

We now form the function

and set

D(A) = *(b, A) + 84'(b, A).

D(X) is a known function of one variable A, and every value of A for which

is obviously an eigenvalue of our problem, since for such values of )4 (3.1) and (3.3) are satisfied simultaneously, i.e., both of the conditions (1.9) are met. Thus, the problem of the existence of eigenvalues reduces to a study of the roots of D(A). By using this fact, it can be shown that the problem has an infinite set of real eigenvalues (see Sec. 4), which can be written as a sequence of the form

&, < hl < * * * < A,, < * * * ,

where

4. Eigenfunctions and Their Orthogonality

Let A be an eigenvalue of our boundary value problem, and let @(x) be an eigenfunction corresponding to A. Then, it is easy to see that every function of the form C@(x), where C is an arbitrary nonzero constant, is also an eigen- function corresponding to A. We shall not consider such linearly dependent eigenfunctions to be different, and in fact any of them can be regarded as a "representative" of the family of functions of the form where C # 0.

We now ask whether two linearly independent eigenfunctions @(x) and Y(x) can belong to the same eigenvalue.3 With our hypotheses, the answer

3 If aQ)(x) + bY(x) r 0 implies a = b = 0, then @(x) and Y ( x ) are said to be linearly indkpendent. Cf. Ch. 2, Rob. 9. (Translator)

Page 266: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

252 THE EIOENPUNCTION METHOD AND ITS APPLICATIONS CHAP. 9

is negative. In fact, suppose that @(x) and Y ( x ) belong to the same eigen- value. Then, by a familiar property of linearly independent solutions of a differential equation, we would have

everywhere on [a, b], and in particular at x = a.4 But according to the first of the conditions (log), we have

a@(a) + p' (a) = 0, a'Y(a) + pY'(a) = 0,

which by (4.1) would imply thqt a = 0, p = 0, thereby contradicting the hypothesis (1.5). Thus, to within a constant factor, only one eigenfunction correspondp to each eigenvalue.

LEMMA 1. Let

where @ is a function dependhg on x. ( I f O also depends on other variables, e.g., on t, we write the partial *derivative with respect to x.) men, the identity

holds for any twice differentiable fictions @ and 'Y.

The proof is an immediate consequence of replacing L(@) and by their expressions as given by (4.2).

LEMMA 2. if d and Y sathfy the boundory conditions (log), then

[@y' - @'Y'L, = [dy' - 8'\Y], 0. (4.4)

Proof. The numbers a and p, which do not vanish simultaneously according to (1.5), satisfy the homogeneous system

This is possible only if the determinant of the system vanishes, i.e.,

a b 4 The symbol 1 I denotes the determinant ad- bc. The functional determinant

c d (4.1) is usually called a Wronskian. (ZPanslator)

Page 267: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 4 THE BIOENFUNCTION METHOD AND FIS APPLICATIONS 253

The second part of equation (4.4) is proved in the same way.

We now show that any two eigenfunctions @(x) and Y(x) correspond- ing to different eigenvalues A and p, respectively, are orthogonal on [a, b], with weight r.

Proof. Let 8 and \Y satisfy the equations

with the same boundary conditions (1.9). We multiply the first equation by Y and the second by (O and then subtract the first equation from the second. By Lemma 1, we obtain

Therefore we have

By Lemma 2 x=b [p(@Y' - (08Y], = 0,

so that (4.5) implies that

Since h # p, it follows that

as was to be shown.

Remark. In Sec. 3, we said that the eigenvalues are real, but we did not give a proof. The reality of the eigenvalues can be deduced from the orthogonality of the eigenfunctions just proved. In fact, if h = p + iu (V # 0) is an eigenvalue and if 8 ( x ) = cp(x) + i+(x) is the eigenfunction corresponding to A, then substituting in (2.1), we obtain

But then we also have the equation

which means that X = p - iv is also an eigenvalue and that the function

Page 268: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

254 THE EIGENFUN~ON METHOD AND ITS APPLICATIONS CHAP. 9

$(x) = p(x) - i+(x) is the eigenfunction corresponding to X. But this implies that

which is impossible, for as just proved, O and must be orthogonal since h # X.

5. Sign of the Eigenvalues

The following theorem gives more precise information about the e i p values for the case where q ( 0 for a ( x ( b, a situation which is encountered very often in the applications.

THEOREM. I f r > 0, q 4 0 and if the boundary conditions imply that

then all the eigenvalues of the boundary value problem for equation (2.1) are nonnegative.

Proof. Let A be an eigenvalue and let @(x) be the corresponding eigenfunction. Multiplying (2.1) by 8(x) and integrating, we obtain

(p@')'O dx + Ib q@z dx = , A Ib r@2 dx, a a

whence, integrating by parts :

~ = b - jb p@t2 dx + Ib q@2 dx = - A Ib r@2 dx. (5.2) [P@@ lxua a a a

It follows from (5.1) and the condition q ( 0 that the left-hand side of (5.2) is < 0. Therefore A >, 0, and moreover, h = 0 is possible only if q = 0,8' r 0, i.e., only if the equation (2.1) has the form

and the function @ = const is an eigenfunction.

Remark. The condition (5.1) seems to be quite artificial, but actually it is not. In fact, it is satisfied for the very boundary conditions most often encountered in the applications, namely

Page 269: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 5 THE EIGENFUNCTION METHOD AND ITS APPLICATIONS 255

where h and H are nonnegative constants. This is obvious for the first and second cases. In the third case,

[ p 4 @ ' ~ ~ ~ = - Hp(b)W(b) - hp(a)O2(a) ( 0,

since @'(a) = h@(a), cP'(b) = - H@(b).

6. Fourier Series with Respect to the Eigenfunctions

t A , . . . A , . . . be all the eigenvalues of our boundary value problem, arranged in increasing order, and let

be the corresponding eigenfunctions, which for simplicity we regard as having the normalization

Then, for every function f(x) which is absolutely integrable on [a, b], we can form the Fourier series

where

and the following propositions, which we cite without proof, are valid:

THEOREM 1. Iff (x) iS continuous on [a, b], ff f (x) has a piecewise smooth (but possibly discontinuous) derivative, and f (x) sati$,fies the boundary conditions of our boundary value problem, i.e.,

then the Fourier series of f(x) with respect to the eigenfunctfohp converges to f(x) absolutely and unlfomly.

The conditions (6.4) may appear artificial, but if we recall how our problem arose [see the first formula in (1.4) and (1.14)] and regard f(x) as the initial value of the function u(x, t) , writing f(x) = u(x, 0), then for t = 0, the conditions (1.3) become just the conditions (6.4).

THEOREM 2. If the function f (x) is piecewise smooth on [a, b] (but either continuous or discontinuous), then the Fourier series of f(x) with

Page 270: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

256 THB B I O ~ ~ ~ ~ O N ~ O D AND m APPLICATIONS CHAP. 9

respect to the eigentctions converges for a < x < b to the value f(x) at every point of continuity and to the wZue

at every point of ciiscontinuity.

Instead of the system (6.1), which is orthogonal with weight r, we can consider the system

which is orthogonal in the ordinary sense, and for which

Let 16 f(x) be a square integrable function. Then, with respect to the new system, its Fourier coefficients have the form

i.e., they agree with the Fourier coefficients off(%) with respect to the system (6.1).

Applied to the function t/;f(x), the completeness condition for the system (6.5) becomes [see equation (7.1) of Ch. 21

If this equation holds for any square integrable function f(x), then instead of reducing the problem to the system (6.5), we simply say that the system (6.1) is complete with weight r. We now prove that this is actually the case.

It is clear from what has just been said that it is sufficient to prove that the ordinary orthogonal system (6.5) is complete. Now, any continuous function @(x) can be approximated in the mean to any degree of accuracy by a function g(x) (with two continuous derivatives) satisfying the boundary conditions of our boundary value problem. [For example, we can choose for g(x) a function such that g(a) = g'(a) = g(b) = gt(b) = 0.1 We shall not give a detailed proof of this fact, which is quite clear from geometric considerations.

Thus, let

where E > 0 is arbitrarily small. According to Theorem 1, the Fourier series of g(x) with respect to the system (6.1) converges uniformly to g(x).

Page 271: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

This means that there exists a linear combination

on(x) = Y O @ O ( ~ ) + Y l@l(x) + + yn@n(x) (6.8)

such that

E I ~ ( X ) - an(x)~ Jyb - g (a 6 x < b).

This implies that

By the elementary inequality

( A + B)* r 2(A2 + B2),

it follows from (6.7) and (6.9) that

This proves that any continuous function can be approximated in the mean to any degree of accuracy by M expression of the form (6.8).

Now let F(x) be any continuous function. Then the function F(X)/-) is also continuous. If we write

h = max r(x), a 6 x 6 b

then by what has just been proved, there exists a linear combination an(x) of the functions (6.1) for which

Therefore, we have

But then the function fi an(x) is a linear combination of functions of the system (6.5), and hence this system satisfies the completeness criterion for ordinary orthogonal systems (see Ch. 2, Sec. 9), i.e., the system (6.5) is complete, as was to be shown. Thus, we have proved

THEOREM 3. The system (6.1) is complete with weight r, i.e., the relation (6.6) holdr for every square integrable fmction f(x).

Page 272: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

258 THE EIGENFUNCIION METHOD AND ITS APPLICATIONS CHAP. 9

The following result is an easy consequence of Theorem 3:

T H E o m 4. Let f (x) be any square integrable function. Then

where the cn are the Fourier coeflcients of the function )c(x) with respect to the system (6.1).

In other words, the Fourier mies off (x) always converges in the mean to f(x), with weight r. To prove this, we need only apply equation (7.3) of Ch. 2 to the function .\/; f(x) and to the system (6.5).

THEORI&I 5. Any contfnuous function f(x) which is orthogonal with weight r to d the function of the system (6.1) must be identically zero.

In fad, if f(x) is orthogonal to all the functions of the system (6.1), then all its Fourier coefficients are zero. But then by (6.6)

Jab rfqx) dx = 0,

whence f(x) = 0.

The theorem on the expansion of a function in series with respect to the eigenfunctions of a boundary value problem (under quite broad hypotheses), as well as the related completeness theorem and its consequences, were first proved by the prominent mathematician V. A. Steklov.5

7. Does the Eigenfunction Method Always Lead to a Solution of the Problem?

The eigenfunction method will certainly lead to a solution of the problem posed in Sac. 1 if first of all, the functionsflx) and g(x) defined by (1.4) have series expansions (1.14), with respect to the eigenfunctions @.(x), which converge to f (x) and g(x), and if secondly, the coefficients A, and Bn defined by (1.16) are such as to guarantee the convergence of the series (1.12) and justify differentiating it twice term by term. However, whether or not these conditions are met, every time the problem has a solution, the solution can be found bt the form of a series (1.12) by the method indicated in Sec. 1. This implies that the solution is unique, a fact which can often be deduced by an argument based on the physical content of the problem. This physical content also allows us to decide whether the problem has a solution in the

The theory given in this chapter is usually associated with the names of the mathema- ticiana J. C, F. Sturm and J. Liouville. (Translator)

Page 273: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

first place. In view of what has just been said, this explains why the physicist or engineer, by using the eigenfunction method and manipulating the series as if term by term differentiation and other operations were justified, even when in fact this is not the case, nevertheless arrives at correct results.

A more precise formulation of these remarks goes as follows:

' I ~ ~ R B M . Let the fundion u(x, t ) be a continuous solution of the equation (1.1) m the region a 4 x 4 b, t 3 0, satlsfhg the bounthry conditions (1.3) and the initial conditions (1.4). Thon

where the @,,(x) are the eigenfunctions associated with the bowr&ay value probZem.6 The functions Tn(t) can be found from the equation

subject to the initial conditions

where the qumrtities C. and c,, are the Fourier coeflcients of f(x) and g(x) [cf. the initial conditions (1.4)J with respect to the system of eigen- fimctions @,,(x). (It is assumed that the derivatives au/at and azu/at2

are continuous and bounded in every region of the form a < x c b, 0 < t < to.)

Proof. We multiply equation (1.1) by the function

(see Sec. 2). Then according to (2.2) and (23), we obtain

By (4.2), this can be written as

and instead of (2.1) we can write U@)= -m.

6 For simplicity, we assume that the eigenflrnctions are normalized as in (6.2).

Page 274: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

260 THE EIGENFUNCTION METHOD AND ITS APPLICATIONS CHAP. 9

Therefore, the relation

W J = - h # n (n = 0, l ,&. . .) (7.4)

holds for the eigenfunctions of the boundary value problem. By the hypothesis of the theorem and by Theorem 2 of Sec. 6, for

a < x < b and every t r 0, the function u(x, t) can be expanded in a series of the form (7.1), where

It follows from (7.4) that

so that

Tdt) = - J" u(x, t)L(On) dx, An a

The last term vanishes because of Lemma 2 of Sec. 4. Thus we have

whence, using (7.3), we obtain

On the other hand, differentiating (7.5) twice with respect to t gives

where the differentiation behind the integral sign is legitimate because of our hypotheses concerning &/at and a%/aP. Comparing (7.7) and (7.8), we obtain (7.2). Furthermore, since u(x, t) is continuous in the region a ( x ( b, t 2 0 and since lim u(x, t) = f(x), it follows from (7.5) that t 4

Page 275: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

where the C' are the Fourier coefficients of the function f (x). Since the function Tn(t) is continuous, this is equivalent to the relations

Similarly, we show that

where the Cn are the Fourier coefficients of the function g(x). This completes the proof of the theorem.

Thus, if the problem in which we are interested has a solution at all, the solution can be found in the form of a series (1.12) by the method of Sec. 1. On the other hand, quite often this method leads to a function u(x, t) which does not have a derivative everywhere. Spch a u(x, t) cannot be regarded as a solution of the problem in the exact sense of the word, since u(x, t) certainly must satisfy the differential equation! However, because of the theorem just proved, in this case it is useless to look for an exact solution, since if such existed, it would have to coincide with u(x, t). This compels us to be satisfied with the function u(x, t) that we have already found; we shall call this function a generalized solution of the problem. It can be shown that with our hypotheses, the series (1.12) obtained by the method of Sec. 1 always defines a function to which it converges either in the ordinary sense or in the mean, and therefore the eigenfunction method always gives a general- ized solution, if it fails to give an exact solution.

8. The Generalized Solution

What is the practical value of the generalized solution described above? Does it represent anything of use to the physicist or engineer, or is it of purely mathematical interest? The generalized solution is in fact valuable, as will emerge from the following theorem :

THEOREM. Let

be either the exact or the generalized solution of equation (1. l), subject to the conditions (1.3) and (1.4). If

= lim [ r[g(x) - g,,,(x)]2 dX = 0, (8.1) m+ ao

Page 276: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

262 THE EIGENFUNCTION METHOD AND ITS APPLICATIONS CHAP. 9

i.e., if fm(x) and g,(x) converge in the mean (with weight r ) to Ax) and g(x), respectively, as m + 00,' and if

is either the exact or the generalized solution of the equation (1. I ) , subject to the boundary conditions (1.3) and the hit ial conditions

then um(x, t ) converges to u(x, t) in the mean as m + a.

Proof. We recall that

where Cn, C , C,, cmn are the Fourier coefficients of the functions f(x), 6(x), fm(x), gm(x), respe*ively. Since

and

lim A,, = +oo

(see Sec. 3), only a finite number of the X,, can be nemtive. Let A,, ( 0 for n ( N and An > 0 for n > N. Then by (8.2), we have*

Tn = cn cos fin t + -EL sin 1 / ~ t (n > N), f i n

7 In particular, this is the case if f d x ) -+ f(x) and gm(x)+ g(x) uniformly. 8 It may happen that AN = 0, in which case we have

TN = CN + CNZ.

Page 277: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

and similarly

Tmn = C,, cos dG t + % sin 4% t (n > N).

Now consider

(see Theorem 3 of Sec. 6). In view of (8.1), these formulas imply that

lim 2 (Cn - CmJ2 = 0, m-*- n=O

whence

Then, by (8.3), (8.4) and (8.6)

lim [Tn - Tmn] = 4 -00

where

[T,, - Tmn]2 = [ (en - Cmn) cos 4% t + -5" sin fin t] d~n

(8 -8) a 2 [(c. - Cmn)2 + (E" - e m n ) l

f i n

for n > N. All that remains is to consider the relation

Page 278: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

264 ?HB BIOENFUNCI'ION METHOD AND ITS APPLICATIONS CHAP. 9

(we Theorem 3 of Sec. 6). By (Bas), (8.7), and (8.8)

which means that the function urn(%, t) converges in the mean to u(x, t) as m + a. This proves the theorem.

What we have just proved can be summarized as follows: Iff,@) is close to f(x) and gm(x) is close to g(x) (In the sense of autI/orm closeness or ma, clareness in the mean), then the fiurction urn(%, t) is close to u(x, t) in the mem.

We now observe that in actual~problems of physics and engineering, the functions f(x) and g(x) are in general not exact, but rather represent approxi- mations to certain exact functions. Nevertheless, by the theorem just proved, even if the solution of equation (1.1), subject to the conditions (1.3) and (1.4), is not an exact solution but only a generalized solution, it will differ only dighay (in the sense of uniform closeness or closeness in the mean) from the true solution of the problem. This constitutes the practical value of generalized solutions.

Finally, we note one further consequence of the theorem proved above:

4 the firnctions fm(x) and g,(x) are chosen in such a way that the fictions um(x, t) me exact solutibn~ of the correspodhg problems (such fm(x) and grn(x) can always be chosenP), then the exact or general- ized solution of equation (1 .I), subject to the conditions (1.3) and (1 .4), is the limit of the exact S O ~ U ~ ~ O I ~ S &(x, t) fm(x) + f (x) and gm(x) -+ g(x) either uniformly or in the mean.

It follows immediately that the generalized solution is unique.

9. The Inhomogeneous Problem

Instead of equation [1.1), consider the more general equation

subject to the same boundary and initial conditions (1.3) and (1.4). In vibration problems, equation Q~1)'comsponds to the case of forced vibra- tions, while equation (1.1) corresponds to the case of free vibrations. When multiplied by the function

9 For example, we can choose the functions fm(x) and gdx) to be the mth partial sums of the Fourier series off (x) and g(x).

Page 279: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 9 THE EIGENPUNCTION METHOD AND IT3 APPLICATIONS 265

(see Sec. 2), (9.1) takes the form

Suppose now that the boundary value problem corresponding to equation (9.1) has a solution and that F(x, t ) has a series expansion in terms of the eigenfunctions of the boundary value problem for the equation

L(@) = -m (~ee SCC. 2). Then, for t > 0, we write u(x, t ) as the series

where

which is possible because of Theorem 2 of Sec. 6. Repeating the argument given in the proof of the theorem of Sec. 7, we obtain

[see (7.6)] or

where we have used (9.2). Assuming that au/at and Pu/at2 are bounded, we find after differentiating (9.4) twice with respect to t that

Finally, if we set

then (9.5) gives

Page 280: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

266 THE E I G E N F U N ~ O N METHOD AND ITS APPLICATIONS CHAP. 9

Then, proceeding as in the proof of the theorem of Sec. 7 [see (7.9) et seq.], we find

Tn(0) = C,, T(0) = c, (n = 0,1,2,. . .), (9.8)

where Cn and c, are the Fourier coefficients of the functions f(x) and g(x), respectively. Thus, if a solution of the problem exists, it is given by the series (9.3), where Tn is determined from the equation (93 , subject to the conditions (9.8).

It is remarkable that just as in the caw of equation (1.1), we can amve at the wries (9.3) by carrying out formal operations with series, without regard to whether or not these operations a n legitimate. In fact, if in equation (9.2) we substitute the series (9.3) and the series (9.6) for F(x, I), differentiate (9.3) term by term, and then equate the coefficients of On(x) to zero, we obtain (9.7). Moreover, if we set t = 0 in (9.3) and require that

we obtain the first of the equations (9.8). If we differentiate (9.3) term by term and again set t = 0, we find that

which is the second of the equations (9.8). Just as in the case of equation (1 .l) in Sec. 7, we can introduce the con-

cept of a generalized solution for equation (9.1). Then it turns out that for any continuous F(x, t), the series (9.3) d e w a function u(x, t) to which it converges either in the ordinary sense or in the mean, so that the problem always has either an exact or a generalized solution. Moreover, this solution is unique, since if U and V are two solutions of the boundary value problem for equation (9.1), the ftnction u = U - V is also a solution of the boundary value problem for equation (1.1) with zero initial conditions. But as ob- served in Sea. 7 and 8, the boundary value problem for equation (1 .l) has a unique solution (exact or generalized). Since the function which is identically zero is a solution of the boundary value problem for equation (1.1) with zero initial conditions, it follows that u r 0, i.e., that U r V. As concerns the practical value of generalized solutions, we can repeat the considerations of Sec. 8.

10. Supplementary Remarks

We considered equation (1.1) only for the sake of being explicit. All our considerations are also applicable to equation (1.2), with the same

Page 281: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

boundary conditions and with the initial condition

In this case, the method of Sec. 1 gives

where the quantities Tn are found from the first order differential equation

where Cn (n = 0, 1.2,. . .) are the Fourier coefficients of f(x). The theorem of Sec. 7 needs only a corresponding change in its statement,

i.c, one has to consider (10.3) instead of (7.2) and require the continuity and boundedness of only &/at. We suggest that the reader prove the new version of the theorem as an exercise.

The concept of a generalized solution can also be introduced in the case of equation (1.2). But this case is essentially different from the case of equation (1.1), in that the generalized solution turns out to be the exact solution as well. This results from the fact that all but a finite number of the A,, are positive, and for such A,,, (10.3) gives

Consequently, it turns out that the series (10.2) is convergent and can be differentiated term by term any number of thes. i.e., (10.2) gives the exact solution.

The considerations of Sec. 9 are also applicable to the inhomogeneous equation

with the same boundary and initial conditions (1.3) and (10.1). The solution Is given by the series

where Tn is now defined by the equations

In view of the results of Secs. 7-10, we shall not worry about the con- vergence of series in the problems to be solved in Part 11, since we now know that the series obtained as a result of solving such a problem h a y s defines a solution (exact or generalized). It is true that strictly speaking, some of the problems in Part 11 do not fall into the category already studied, but it can be shown that the previous considerations apply to them as well.

Page 282: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

268 THE EIOENFUNCTION METHOD AND ITS APPLICATIONS CHAP. 9

Part 11. APPLICATIONS

I I. Equation' of a Vibrating String

Consider a stretched homogeneous string, fastened at both ends, and suppose that the equilibrium position of the string is a straight line. Take this straight line to be the x-axis, and let the ends of the string be located at the points x = 0 and x = I, where I is the length of the string. If the string is displaced from its equilibrium position (or if certain velocities are imparted to the points of the string), and the string is then released, it begins to vibrate. We shall consider only the case of small vibrations; then, the length of the string can be regarded as unchanged. Moreover, we shall regard the vibrations as taking place in one plane, in such a way that each point of the string moves in the direction perpendicular to the x-axis.

Let u(x, t) be the displacement at time t of the point of the string with abscissa x. Then, for every fixed value of t, the graph of the function u(x, t ) obviously represents the form of the string at the time t. The element AB of the string (see Fig. 47) is acted upon by the tension forces TI

and T2, which are directed along the tangent to the string. For the time being, we assume that no other forces act on the string. In the equilibrium position, the tension T is the same at all points of the string. To the extent that we can assume that the length of the string does not change (see above remark), we can also assume that the tension in the string does not change. Therefore, TI and T2 have the same magnitude as T, although they have different directions, and because of the curvature of the element AB, one direction is not quite the negative of the other. Hence, the force acting on the element AB in the direction of the u-axis is

T[sin (9 + Acp) - sin cp]

o Tpan(cp +hip) - tanp] = T [ ~ U ( X + Ax, t) - ax ax "I

Page 283: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

where z denotes approximate equality. Regarding the element Ax as being very small, and using Newton's second law of motion, we can write

a2u azu PAX- = T-AX, at2 8x2

where p is the linear density of the string (i.e., the mass per unit length). Setting a2 = T/p and dividing by AX, we obtain

which is the equation for free vibrations of the string. Next, suppose that in addition to the tension T, the string is acted upon

by a force of amount F(x, t) per unit length of the string. Then, instead of equation (1 1. I), we obtain

which is the equation for forced vibrations of the string. We now study the following problem: Given the form of the string and

the velocity of its points at the initial time t = 0, what is its form at the arbitrary time t ? Mathematically, this problem reduces to solving equation (1 1.2) in the case of free vibrations, and equation (1 1.3) in the case of forced vibrations, subject to the boundary conditions

and the initial conditions

where f(x) and g(x) are given continuous functions, which vanish for x = 0 and x = 1. Equations (1 1.2) and (1 1.3) are special cases of the equations studied in Part I.

12. Free Vibrations of a String

Instead of starting from the formulas already found in Part I, we shall once more go through the derivation given in Sec. 1. We are looking for a solution (different from u = 0) of the form

Page 284: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

270 THE BIOENFUNCT~ON METHOD AND ITS APPLICATIONS CHAP. 9

which satisfies the boundary conditions. Substituting (1 2.1) in (1 1.2) gives

whence (9" T" -3 - -

@ a2T - -A = const,

so that @" = -A@,

If a function of the form (12.1), which is not identically zero, is to satisfy the conditions (1 1.4), then obviously the condition

must be met. Thus, we obtain a boundary value problem for equation (1 2.2), subject to the condition (12.4). In view of the theorem of Sec. 5 (see also the remark made there), all the eigenvalues of our problem are positive.10 Therefore, it is permissible to write A2 instead of A. Then, equations (12.2) and (12.3) take the form

The solution of (12.5) is

0 = Cl cos Ax + C2 sin hx (Cl = const, C2 = const),

where for x = 0 and x = I we must have

@(O) = Cl = 0, @(l) = C2 sin M = 0.

Assuming that C2 # 0, since otherwise @ would be identically zero, we find that M = m, where n is an integer. Setting C2 = 1 gives

7en A,, = (n = 1,2,. . .),

and the corresponding eigenfunctions are

mx @Ax) = sin - I (n = 1,2,. . .).

10 This can also be verified directly by examining the solution of equation (12.2) for X 6 0. By doing this, the reader can convince himself that the condition (12.4) cannot be saMed.

Page 285: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 12 THE EIGENFUNCTION METHOD AND ITS APPLICATIONS 27 1

We do not consider negative values of n, sin= they give the same eigenfunc- tions (up to a constant factor) as the corresponding positive values of n. Thus, in the sense indicated in Sec. 4, only one eigenfunction corresponds to each value of 12.

For X = )in, equation (12.6) gives

Tn = A, cos Ant + Bn sin ah,t a n t mnt = A, cos - I + B, sin - I (n = 1,2,. . .),

so that annt u~(x, t) = (A , MS - I + B, sin -

Thus, to solve our problem, we set

mnt = 2 (A. cos - + B, sin - I I sin 7 n 5 I ) Irnx

and require that11 00 nnx U(X, 0) = 2 An sin - = f(x),

n = 1 I

mn mnt m n at I 1 + B n 7 1 1-0

sin - cos s) sin y] n= 1

cO mn m x = 2 B n T sin = g(x).

n= 1

Therefore, we have to expand f(x) and g(x) in Fourier series with respect to the system {sin (nnx/l)). The formulas for the Fourier coefficients give

i.e., the solrition of our problem is given by the series (12.8), where A, and Bn are determined from formulas (12.9) and (12.10).

11 As in the method of Sec. 1, we differentiate the series term by term.

Page 286: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

272 THE EIO-CTION METHOD AND m APPLICATIONS CHAP. 9

Thus, we see that the vibrational motion of the string is a superposition of separate harmonic vibrations of the form (12.7), or of the equivalsnt form

m t ftltx. a = H, sin (T + a,,) sin --. I

where

The amplitude of the vibration of the point with coordinate x is

and is independent of t. The points for which x = O,l/n, 22/12, . . . , (n - l)l/n, I, remain fixed during the motion, and are known as nodes. Hence, a string whose vibrations are described by formula (12.7) is divided into n segments, the end points of which do not vibrate. Moreover, in adjacent segments, the displacements of the string have opposite signs, and the midpoints of the segments, the so-called antinodes, vibrate with the largest amplitudes. This whole phenomenon is known as a standing wave.

Figure 48 shows consecutive positions of a string whose vibrations are described by formula (12.7). where n = 1,2,3,4. In the general case, where

the vibrations of the string are described by formula (12.8), the funclbmental (mode) corresponds to the component ul with frequency

and period

Page 287: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 12 THE EIGENFUNCTION METHOD AND ITS APPLICATIONS 273

The other vibrational modes of the string, i.e., the overtones, with frequencies

and periods

characterize the timbre or "color" of the sound. If the string is held fixed at its midpoint, then clearly the even overtones of the string, for which the midpoint of the string is a node, are preserved. However, the fundamental and the odd overtones are immediately extinguished, since by holding the midpoint of the string fixed we essentially go from a string of length I to a string of length 112, and changing I to 112 in (12.8) leads to a series containing only even components. Then the overtone with period = 27c/w2 = 4 2 plays the role of the fundamental.

13. Forced Vibrations of a String

Next we consider the case of a periodic perturbing force, i.e., we write

F(x, 1) = A sin at. e

Then

F(x, t) a0

SUlX = A sin a t = 2 Fn(t) sin 7

P n= 1

where

2 1 'IVUC Fn(t) = 7 lo A sin ~t &I - dx I

2A =- [l - (-I)"] sin wt (n = 1,2, • • .). xn

If we write PO w x

u(x, t) = 2 T,(t) sin 7 n= 1

substitute (1 3.2) and (13.1) in (1 1.3). and carry out the required term by term differentiation, we obtain

2A Tm - [l - ( - l r ] sinwt

n=1

Page 288: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

whence a2x2n2

T: + 7 2A Tn - ;;;; [l - (- l)"] sin ot = 0. (1 3.3)

Writing mn a,, = -

I (?a = 1,2,. . .) for simplicity (these will be recognized as the frequencies of the free or characteristic vibrations of the string), we can rewrite equation (13.3) as

24 T: + 03. = - [l - (- ly] sin at. xn Solving this equation, we obtain

Tn = A, cos %t + B,, sin %t + 2A[1 - (- l)"l sin at, xn(a: - 09 (13.5)

if a, # a. To satidy the conditions (1 1.4) and (1 1.5), we require that 00 nnx

u(xt 0) = 2 Tn(0) sin = f (x), n= 1

Wx, 0) 00 mx

at = 2 TL(0) sin = g(x). n = l

A calculation of the Fourier coefficients of f(x) and g(x) gives

2 1 m x Tn(0) = A,, = 7 f(x) sin - dx, 1

2 1 ltnx = I0 B(X) sin

B~ = 2- 1' g(x) sin m x - dx - 2A4l - (- l)"l l o , o 1 nnon(oi - 01)

where we have used (13.5). Substituting (13.6) and (13.7) in (13.9, and then substituting the resulting expression for Tn in (13.2), we obtain

00 mx U(X, t) = 2 (An cos ant + En sin ant) sin -

n = l I 4A " sin [(2k + l)rtx/I] + - sin ot 2 7F k=o (2k + 1)(4+1 - a2) (1 3.8)

- 4&-$ sin o,+,t sin [(2k + l)n:x/I] = LPO (a + 1)%+1(4k+l - a2)'

Page 289: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

where we have written

m x iin = 1 j" g(x) sin --r h. la, 0

Recalling the expression for an, the reader will easily recognize that the first sum in the right-hand side of (13.8) is the function giving the free vibrations of the string, subject to the conditions (1 1.4) and (1 1.5). [See (12.8), (12.9). and (12. lo).] Therefore, the second and third sums give the "correction" caused by the presence of the perturbing force. The second term represents what is sometimes referred to as the "pure" forced vibrations, since they occur with the frequency of the perturbing force.

Equation (13.5) holds if on # o. We now examine the situation when on = a, i.e., when the frequency of the perturbing force is the same as one of the characteristic frequencies of the string. Then, equation (13.4) gives

At Tn = A,cosat + &sinat - - [l - (-I)"] cos ot. ma

This shows that when n is odd, the amplitude of the nth vibration in the term Tn(t) sin (mxll) of the sum (13.2) is

which becomes unbounded as t increases. In this case, we say that resonance occurs.

14. Equation of the Longitudinal Vibrations of a Rod

Consider a homogeneous rod of length I. If the rod is stretched or compressed along its longitudinal axis, and then released, it executes longi- tudinal vibrations. We choose the x-axis along the axis of the rod, and we assume that in the equilibrium position, the ends of the rod are located at the points x = 0 and x = 4 1 . Let x be the abscissa of a cross section of the rod at rest, and let u(x, t) be the displacement of this cross section at the time t. Consider an element AB of the rod, whose ends are located at the points x and x + Ax when the rod is at rest, and let the ends of this element have the new positions A' and B' at the time t, with abscissas x + u(x, t) and x + Ax + U(X + AX, I), respectively (see Fig. 49). Thus, at the time t, the element AB has length Ax + u(x + Ax, t) - u(x, t), i.e., its absolute increase in length is

u(x + Ax, 1) -- u(x, 1) = "(x + AX, t) hx (0 < 0 < I), ax

Page 290: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

CHAP. 9

while its relative increase in length is

In the limit as Ax + 0, the relative increase in length of the cross section of the rod at x is

According to Hooke's law, the force T acting on the cross section at x is given by the formula

when E is the modulus of elasticity (Young's modulus) of the material from which the rod is made, and s is the cross-sectional area of the rod.

Returning to the element AB which has the new position A'B' at time t, we note that AB is acted upon by forces TI and T2, applied at the points A' and B' and directed along the x-axis. (For the time being, we consider no other forces.) The resultant of these forces is

au(x + Ax, 1) q x , 1) T2 - TI = ( ax - ax )

and is also directed along the x-axis. Regarding the element AB as being very small, we can write

where p is the density of the material from which the rod is made. Setting E/p = a2 and dividing by Ax and s, we obtain

which is the equation forfree vibrations of the rod. This equation has the same form as the equation for free vibrations of a string.

Page 291: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

If the rod is acted upon by an additional force of amount F(x, t) per unit volume, then instead of (14.1), we have

so that

This is the equation for forced oscillations of the rod [cf. (1 1.3)]. We now pose the problem of finding the displacement of the cross sections

of the rod at any time t, with specified initial and boundary conditions. The boundary conditions can take various forms:

Case 1) The rod is fastened at both ends:

Case 2) One end of the rod is fastened, and the other end is free:

(the force on the free end of the rod is zero and hence aulax = 0) ; C w 3) Both ends are free. We shall devote our attention to Case 2, where the boundary conditions

(14.4) are met. The initial conditions have the familiar form

i.e., the initial displacements and initial velocities of the cross sections of the rod are specified.

15. Free Vibrations of a Rod

As in the case of the vibrating string (see Sec. 12), we look for particular solutions of the form

u(x, t) = @(x)T(t),

and amve at the equations

with the boundary conditions

Q(0) = @'(I) = 0.

Page 292: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

278 THE E I G E N M I N ~ O N METHOD AND m APPLICATIONS CHAP. 9

Equation (15.1) has the solution

@(x) = Cl cos hx + C2 sin hx (Cl = const, C2 = const).

According to (1 5.3). for x = 0 and x = 1 we must have

@(0) = Cl = 0, @'(Z) = C2A cos M = 0.

Assuming that C2 # 0, since otherwise @(x) would be identically zero, we find that M = (2n + 1)x/2, where n is an integer. We write

n,= (2n + 1) sc 21 (n = 0, 1.2 ,... ),

(1 5.4)

Qn(x) = sin A,,x = sin (h + 1)rvc 21

(n = 0, 1, 2 ,... ). (The negative values of n give no new eigenfunctions.) For h = A , equation (1 5.2) leads to

Tn(t) = A, cos ah,/t + B,/ sin aAnt (n = 0, 1, 2, . . .), and therefore

un(x, t ) = (A, cos Ant + Bn sin a)k?) sin A s (n = 0, 1,2, . . .). To solve our problem, we form the series

00 00

u(X9 t' = 2 un(X, t ) = 2 (An cos .Ant + Bn sin =Ant) sin A s , (15.5) at ,PO n=o

and require that a0

U(X, 0) = 2 A, sin )Uc = f (x), n=O

W x , 0) at = [% (- AgAn sin aAnt + Bna& cos abt) sin hx I t=o

00

= 2 Bph. sin A,,x = g(x). nu0

A calculation of the Fourier coefficients of f(x) and g(x) with respect to the system {sin )Uc) gives

jO1 f(x) sin ~p A, = (n = O,l,2 ,... ),

sin2 dx

f ~ ( x ) sin dx

= ( I sin2 dx (n = 0, 1,2,. . .).

0

Page 293: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 15

But

and therefore

2 1 A n = 7 1, f ( ~ ) sin A s d~

B,=- J1 g(x) sin a dx ah1 0

[see (15.4)J. Thus, the solution of our problem is given by formula (15.5), where An

and Bn are determined from (15.6), i.e., the vibrational motion of the rod is a superposition of separate harmonic vibrations

un = (A, cos alnt + Bn sin ahat) sin A s , (1 5.7) or

Un = Hn sin (aht + an) sin A&, where

The amplitude of the vibrational motion described by (15.7) is

which depends only on the position x of the cross section, and not on the time t. As for the ftequency, it is given by

and hence the period is

In the case of a vibrating rod, the fundamental mode is obtained for n = 0; it has amplitude

(A, sin 51,

Page 294: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

280 THE EIGENFUNCTION METHOD AND I'IB APPLICATIONS

frequency

and period

Therefore, the fundamental mode corresponds to a node at the fixed end (x = O), and an antinode at the free end (x = I ) , as shown in Fig. 50.

16. Forced Vibrations of a Rod

Next, we consider the case where the rod is suspended from the end x = 0 and the perturbing force is the force of gravity, i.e.,

where F(x, t) is the force per unit volume, p is the density of the rod, and g is the acceleration due to gravity. In this case, the equation for the vibrations takes the form

[see (14.3)], again subject to the boundary conditions (14.4) and the initial conditions (14.5). We set

QO

U(X, t) = 2 Tn(t) sin XA, n=O

F(x, t) QO

= g = 2 Fn sin A&, P n=O

where

j'i g sin dx Fn = 2g

= K (n = 0, 1,2 ,... ). 1: sin2 tix

Page 295: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 16 THE EIGENFUNCTION METHOD AND ITS APPLICATIONS 281

Substituting the series (16.2) in (16.1) and transposing all the terms to the left, we obtain

sin A,p = 0, n=O

whence

The solutjon of this equation has the form

2g T' = An cos ah,t + Bn sin aAnt + - la2Ai (n = 0, 1.2, . . .).

To satisfy the conditions (14.5), we require that

a0

au(x9 O) = 2 TL(O) sin A!,X = g(x). at n u 0

A calculation of the Fourier coefficients of f(x) and g(x) with respect to the system {sin An) gives

2g 2 l ~ n ( 0 ) = A. + = 7 jo f(x) sin A p d ~ ,

2 1 T ~ ( o ) = B,,uA,, = 7 g(x) sin ~ , , x dx

[cf. (1 5.6)], so that

Therefore, we have

U(X, I ) = 2 ( I n cos aht + Bn sin aAnt) sin A,p n-0

2g 5 cos a&? sin )Uc - - + $5 sin Anx. la2 ,=o A: n=O %!

The reader will recognize at once that the first sum on the right is the function giving the solution of the problem of the free vibrations of a rod with the same

Page 296: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

282 THB BLOWFUNCTION -OD AND m APPUCATIOW CHAP. 9

conditions (14.5). Hence, the second and third terms give the corrections due to the force of gravity.

17. Vibrations of a Rectangular Membrane

By a membrane, we mean an elastic iilm supported by a closed plane cum. When the membrane is at rest, all its points lie in one plane, which we take to be the xy-plane. If the membrane is displaced from its equilibrium position and then released, it begins to vibrate. We consider only small vibrations of the membrane, assuming that the area of the membrane does not change and that each of its points vibrates in a direction perpendicular to the xy-plane. Let u(x, y, t ) denote the displacement at time t of the point (x, y) of the membrane from its equilibrium position. Then, by a derivation similar to that made in the case of the ~ t r h g , one finds that the equation for the free vibrations of the membrane has the form

while the equation for forced vibrations oc the membrane is

where c2 = T/p, T is the tension in the membrane, p is its surface density, and F(5 y, t) is the force per unit area acting on the membrane.

The problem of the vibrating membrane can be posed as follows: Find the solution of equation (17.1) or (17.2), i.e., find the displacement of the points of the membrane at any time t, subject to the condition

met on the (fixed) boundary of the membrane, and the initial conditions

U(X, Y, 0) I f (x, Y) (17.4)

(specifying the initial displacement of the membrane) and

(specwng the initial velocitiq of the points of the membrane). We now study the cam of thefrce vibrations of a membrane in the shape

of a rectangle R: 0 6 x < a, 0 6 y $ b. The problem differs from the problem considered in Part I in that the function u depends on three rather than two variables. Nevertheless, we again apply the eigenfundion method and begin by loddng for particular solutions of the form

Page 297: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 17 THE EIGENFUNCTION METHOD AND ITS APPLICATIONS 283

which are not identically zero and which satisfy the condition (17.3) on the boundary of the rectangle R. Differentiating (17.6) and substituting the result in (17.1), we obtain

Hence

where, as is easily seen, the function O satisfies the condition

@ = O (17.9)

on the boundary of the rectangle. Thus, we now consider equation (17.7), with the boundary condition

(17.9). We fix h and look for particular solutions of (17.7) of the form

which satisfy the condition (17.9) on the boundary. Substituting (17.10) in (17.7) gives

f = - e. + 4J = - k2 = const,

'P

whence

'Pa +k* 3 0 , qJu+z2J I =o, where we have written

12 The fact that this constant should not be chosen to be positive is clear just from the fact that otherwise the coefficient of T in equation (17.8) would be negative, so that the solution of (17.8) would not be periodic, i,e., contrary to experience, we would not have vibrations.

13 If the constant were positive, we could not satisfy the boundary condition.

Page 298: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

284 THE EIGENFIJNCTION METHOD AND ITS APPLICATIONS CHAP. 9

For the solutions of (17.11). we find

p(x) = Cl cos kx + C2 sin kx, $(x) = C3 cos ly + C4 sin ly,

where C,, C2, C3, and C4 are constants. The boundary condition implies that

~ ( 0 ) = ?(a) = 0, +(O) = +(b) = 0,

and hence

Cl = C3 = 0, C2 sin ka = Cq sin lb = 0,

so that ka = mrr, lb = m, where m and n are integers. Setting C2 = C4 = 1 and writing

we find

xmx qm(x) = sin kmx = sin -9 a

( m = 1.2 ,...; n = 1.2 ,...) ZflY $&) - sin Z,,x = sin -.

b

(We do not consideranegative values of m and n, since they give the same functions q,,, and +,, to within a constant factor.) According to (17.10) and (17.121, if

A = )Cnn = = x .\/(m2/&) + (n2/b?, (17.13)

we obtain the following particular solutions of equation (17.7) satisfying the boundary condition :

nmx m y . Y ) = qbn(x)#n@) = sin - sin a

We now solve equation (17.8) for every h = A,,,,,; the result is

Tmn(t) = Am,, cos ~&,,t + BmM sin cXmNt.

Therefore, the functions

m x leny u,&, Y, t ) = (Amn cos cA,,,,,~ + Bmn sin cXmnt) sin - sin a

are particular solutions of (17.1) satisfying the boundary condition (17.3).

Page 299: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

sec. 17 THE EIGENFUNCTION ~ O D AND ITS APPLICATIONS 285

To obtain the solution of equation (17.1) satisfying the initial conditions, we set

a0 lcmx m y 2 (A, cos cA,J + BmH sin chmnt) sin - sin (17.15)

m, n=1 a

and require that

nmx xny u(x, y, 0) = A, sin - a sin 7 = f (x, Y), m. n= 1

(- Am&,,,, sin cX,,t + &&Amn cos cX,,~) sin - a

nmx m y = Bm&Ln sin- sin 70

m,n=l a

Assuming that f(x, y ) and g(x, y) can be expandeg in double Fourier series with respect to the system {sin ( m x l a ) sin (nny/b)) and using the results of Ch. 7, Sec. 4, we obtain

scmx scny = "1 j f ( x , Y ) sin- ab n sin - dx dy a b

and

4 x m m y ~mn~hmn = - ab ~ J ' B ( X , Y ) sin -;r sin 7 tk dy,

nmx nny B m n = J /g(x, Y) sin - sin ak dy. abchm R a

(17.17)

Thus, the solution of our problem is given by the series (17.15), where A, and Bmn are calculated by usinp formulas (17.16) and (17.17).

.-

14 In Ch. 7, Sec. 4, we expanded functions in double Fourier series in a rectangle of the fonn - 1 d x d 1, - h d y d h. If instead, we want to expand a function f (x, y) in a rec- tangle of the form 0 6 x G a, 0 G y d b, with respect to the system (sin (nmxla) sin (xnylb)), this can be done by making the odd extension of f(x, y), first in the variable x and then in the variable y. Then, the formulas for the Fourier coeflicients take the form (17.16).

Page 300: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

286 THE EIGENFUNCTION METHOD AND rrs APPLICATIONS CHAP. 9

The frequencies of the characteristic vibrations of the rectangular membrane have the form

[see (17.13) and (1 7.1411, with the corresponding periods

There is an important difference between the case of the vibrating membrane and that of the vibrating string: In the case of a string, every characteristic frequency corresponds to its own configuration of nodes, whereas in the case of a membrane, the same characteristic frequency can correspond to several configurations of nodal lines (i.e., lines or curves which remain fixed during the course of the motion).

We illustrate this situation, using the particularly simple case of a square membrane, where we set a = b = 1. Then, the frequencies of vibration are

and instead of (17.14), we can write

Umn = HmN sin (omnt + amn) sin zmx sin my,

where

A, sin a,,,,, = - , Bmn cos a,,,,, = -* H * n H,,

For the fundamental mode, i.e., for m = n = 1, we have

a11 = xc f i , ull = fil sin (ollt + all) sin nx sin xy,

from which it is clear that there are no nodal lines at all. Next, we set m = 1, n = 2 or rn = 2, n = 1. Then, the same frequency

corresponds to the two modes

2412 = HI2 sin (wt + a13 sin rur sin by, 242, = Hal sin (at + aZl) sin 2zx sin zy,

the first of which has the nodal line y = f and the second the nodal line x = f. Moreover, the frequency o also corresponds to the "compound"

Page 301: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

mode ulp + ~ 2 1 , which in general leads to a new nodal line. In fact, writing a12 = 6z1 = 0 for simplicity, we obtain

ul2 + 2421 = sin at(H12 sin nx sin 2ny + sin 2xx sin ny),

whence it follows that the points satisfying the equation

HI2 sin xx sin 2ny + H2* sin 27rx sin z y = 0 (1 7.18)

form a nodal line. In particular, for H12 = HZl we obtain

sin ICX sin Zny + sin 2nx sin ny = 2 sin nx sin ny (cos xx + cos ny) = 0,

which gives the nodal line

x + y = l .

Similarly, for HIS = - HZl, we find the nodal line

Thus, if we vary the coefficients HI2 and Hzl in the "compound" tone 2412 + U Z ~ , equation (17.18) leads to new nodal lines, i.e., an infinite number of nodal lines can correspond to a given frequency. In Fig. Sl(a), we show the simple nodal lines just found for the frequency o = n d .

If we set m = 2, n = 2, we obtain the frequency

w, = ncdfi,

corresponding to the unique mode

u 2 ~ = Hz2 sin (out + a2d sin 27m sin 2xy.

In this case, the nodal line is the locus of the points in the plane for which

Page 302: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

288 THB EIOENFUNCLION -OD AND I'IS APPLICATIONS CHAP* 9

either x = + or y = f [sse Fig. 51(b)]. Finally, in Fig. 51(c), we show the simplest nodal lines corresponding to the fnquency

18. Radial Vibrations of a Circular Membrane

Next, we consider a circular membrane of radius I. For convenience, we introduce polar coordinates in the xy-plane, choosing the center of the membrane as the origin. The change of variables

transforms equation (17.1) into

and equation (17.2) into

We shall discuss only the case of free vibrations of a circular membrane, i a , the case described by equation (18.1), assuming first that the initial displacements and initial velocities of the points of the membrane are independent of the angle 0. Then, it is clear that at any time t, the displace- ment is also independent of 0, so that u = u(r, t). In this case, the vibra- tions are said to be radiul, and equation (18.1) reduces to

The boundary condition has the form

u(2, t) = 0,

and the initial conditions have the form

Following our previous method, we look for solutions of equation (18.3) of the form

Page 303: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

which satisfy the condition (18.4). Differentiating this expression and substituting the result in (18.3), we obtain

whence -

so that

Equation (18.6) is the parametric form of Bessel's equation, with index p = 0 (see Ch. 8, Sec. 1 I), and has the general solution

(see Ch. 8, Sec. 6). Since Y&) is unbounded at r = 0, we must set C2 = 0. Moreover, to obtain a solution different from R -= 0, we have to assume that C1 # 0. Then, it follows from the boundary condition (18.4) that

J o ( W = 0,

i.e., p = M must be a zero of the function J,-,(p). Setting Cl = 1, we obtain

where p,, = )kl is the nth zero of the function Jo(p). For X = An, equation (18.7) has the solution

Tn(t) = A, cos ~ & t + B, sin cXnt (n = 1, 2, . . .). Thus, we have finally found particular solutions of (18.3) of the form

ti&, 2) = (A, cos cAnt + Bn sin cAntJo0@~) (n = 1,2, . . .), (1 8.9)

satisfying the boundary condition (1 8.4). To find a solution of (18.3) satisfying both the boundary condition (18.4)

and the initial c o n d i t h (18.5). we write PO

~ ( r , t ) = 2 (A, cos c)kt + B. sin cXnt)JO(h,r), (18.10) n o 1

Page 304: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

CHAP. 9

and require that

wr, 0) OD

at = [ n o 2 1 (--A&& sin c&,t + B d , , ms C~~,$)J~(A,J) ] t-o

A calculation of the Fourier coefficients of f(r) and g(r) with respect to the system {J&)} gives (see Ch. 8, Sec. 24)

Thus, the solution of our problem is given by the series (18.10), where the coefficients An and B,, are determined fiom the formulas (18.1 1) and (18.12).

The separate harmonic vibrations (18.9) which make up the complicated vibrational motion of the membrane, can be represented in the form

un(r, t) = H, sin ( e u + a,,) Jfij),

where

and the characteristic frequencies of the membrane are

Pn On = Cha Cr The amplitude of vibration of the nth mode is

and depends only on r. The nodal lines are obtained from the equation

[see (18.8) 1. If n = 1, there are no nodal lines. If n = 2, a nodal line is

Page 305: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

obtained for r = pll/p2, if n = 3, nodal lines are obtained for r = p1Z/p3, r = &Z/p3, etc. In Fig. 52, we show the nodal lines and corresponding

sections of a membrane whose vibrations are described by (18.9), for n = l ,2,3.

19. Vibrations of a Circular Membrane (General Case)

In the general case, the problem of the free vibrations of a circular membrane of radius 2 reduces to solving the equation

(see beginning of Sec 18), with the boundary condition

u(l, 0, t ) = 0, (19.2)

and with the initial conditions

"(r, 0,O) = f(r, 0)s

We look for particular solutions in the form of a product

"(r, 0, t ) = @(r, font)*

which are not identically zero and which satisfy the boundary condition ( 1 9.2). Substituting this expression in (19.1). we obtain

Page 306: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

CHAP. 9

Therefore

where to satisfy the boundary condition (19.2), we must require that

Thus, we arrive at a boundary value problem for equation (19.4). To solve this problem, we look for particular solutions of (19.4) of the form

which are not identically zero and which satisfy the condition (19.6). Sub- stituting (19.7) in (19.4) gives

1 1 RnF + - R'F + ;j RFn + X2RF = 0, r

Therefore, we have

Equation (19.9) has solutions of the form cos v0 and sin v0. Since if we change 0 by 2n, we come back to the same point of the membrane, the function u, and hence Q, and F, must have period 27t. Therefore, the number v must be an integer, i.e.,

v = n , (n=O, l ,2 ,... ), and the solutions of (19.9) are17

1s We take the constant to be negative, since otherwise the function T would not turn out to be periodic, and we would not have vibrations, contrary to what actually happens.

16 The constant is taken to be negative, since in a problem like this, the function F(8) must be periodic [see (19.9) et seq.].

17 For v = 0, equation (19.9) has another solution of the form CO, which can be dis- carded, since it is not periodic.

Page 307: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

(The negative values of n give the same functions, to within a constant factor.)

Thus, equation (19.8) takes the form

This equation is the parametric form of Bessel's equation, with integral index n, and has the general solution

Since the solution must be bounded, we are compelled to set C2 = 0, because Yn(b) + a, as r -c 0 (see Ch. 8, Sac. 6). If we put Cl = 1, then by the boundary condition (1 9.6)

i.e., M = p must be a zero of the function Jn(p). We now write

where pnm is the mth positive zero (in order of increasing size) of the function Jn(p). Then, the boundary value problem for equation (19.4), subject to the boundary condition (1 9.6). has the eigenvalues )4, [see (19.1 I)], and the eigenfunctions [see (1 9.7), (19.1 O), (19.1 I)]

0) = Jn(Lr) cos nB, (n = 1,2 ,...; m = 1,2 ,... ).

@:rn(r, 0) = Jn(XnJ) sin d

For A = Am, equation (19.5) gives

T,, = Anm cos cAnmt + B,, sin ch,,t.

Therefore, we have the following particular solutions of (19.1), satisfying the boundary conditions (19.2) :

u,,,,,(r, 0, I) = (A, cos cA,t + B,, sin cAnmt)Jn(A,r) 00s n6 ( m = 1,2 ,...; n = 0 , 1 , 2 ,... ),

~:~(r, 0, t) = (A& COS CA,t + B:m sin CA,t)J,,(Anmr) sin no (m = 1,2 ,...; n = 1,2 ,... ).

Page 308: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

To obtain a solution of equation (19.1) which also satisfif the 1.. dn ~ a r y conditions (19.3). we form thif series

~ ( r , 0, t ) = 2 2 [(Anm cos cLmt + Bnm sin cAnmt) cos n=Om=l

+ (A:m cos cLmt + B+, sin c ~ ~ t ) sir, J- J , , : A ~ ~ ) and require that

= 2 2 (Am COs + A 2 sin nO)Jn(h,mt) = f(r , 8). n=Om=l

(19.13)

= { 2 [(- AnMcAm sin eAmt + BmcIm COP c L t ) ms n0 n=Om=l

+ ( - ALcInm sin cXnmt *

+ =cAm cos c I n ) sin n0 ] J~(A,,,,,~)) t-0

+ B:rncXnm sin n0)&@,,,~) = g(r, 0). @

To find the coefficients in these expansions, we now argue as follows: Let

where

= x f(r, 8) sin nB d0

i.e., we expand the function f(r, 0) in a trigonometric Fourier series with respect to the variable 0. We then expand each of the functions fAr) and f t ( r ) in Fourier series with respect to the system {Jn(h,mr)}. The result is

Page 309: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

?L. 19

where

THE ElGENFUNCI'ION METHOD AND ITS APPLICATIONS 295

Substituting (19.16) in (1 9.14) gives

f (r, 0) 2 2 (Cnm cos d + C:m sin nO)Jn@nmr). n = O m = l

Comparing this with the first of the formulas (19.13), and using (19.15) and (19.16). we obtain

- J' dr if g(r, 0) sin n8Jnnmr) dB - J o -=

In just the same way, we find that

BnmChnm = rg(r, 0) cos n0 JR@,,,,,r) do

1' R

B:mclnm = 7Ei2J2 + J_ rg(r, 0) sin n O J n ( k ~ ) n+l Pnm

Thus, the solution of our problem is given by the series (19.12), where the coefficients are determined from formulas (19.17) and (19.18).

There are a great variety of nodal lines corresponding to the simple harmonic vibrations unm and utm, which make up the complex vibration (19.12). For example, for uo,, uo2, and U03, one finds the nodal lines shown in Fig. 52, while for 2412, ql and U ~ B one finds the nodal lines shown in Fig. 53.

Page 310: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

296 THB EIO-CIION METHOD AM) ITS APPLICATIONS CHAP. 9

As in the case of a rectangular membrane, an infinite set of different con- figurations of nodal lines can correspond to the same frequency (depending on the coefficients A,, B , A:, B:J.

20. Equation of Heat Flow in a Rod

Consider a homogeneous cylindrical rod, whose lateral surf= is insulated from the surrounding medium. Choose the x-axis along the axis of the rod, and let u(x, t) denote the temperature at time t of the cross section of the rod with abscissa x. Let AB be the element of the rod lying between x and x + Ax (see Fig. 49). Let At be a time interval which is so small that we can assume that the temperature of the cross sections at x and x + Ax does not change (in time). It has been established experimentally that the amount of heat q flowing through a rod whose ends are held at two constant tempera- tures is proportional to the difference between the temperatures, to the cross- sectional area of the rod and to the time interval At, while q is inversely proportional to the length of the rod. Therefore, the amount of heat flowing through the element AB is

where Kis a constant of proportionality called the thermal comktivity of the material from which the rod is made, and s is the cross-sectional area of the rod. In the limit as Ax-, 0, we obtain the amount of heat Q(x) flowing through the cross section at x in the time At:

au Q(x) = K E s At.

Now, it can easily be seen that the amount of heat AQ which the element AB receives in the time At is

AQ = Q(x + ax) - Q(x)

= K s A t [ au(x + Ax, t) - ax ax

Page 311: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 20 THE EIGENFUNCTION METHOD AND ITS APPLICATIONS 297

(It should be kept in mind that heat flows in the direction opposite to the direction in which the temperature increases.) The quantity AQ can also be calculated by another method: Suppose that the element AB is so small that at any given time the temperature of all its%ross sections can be regarded as being the same. Then

AQ = cps Ax [u(x, t + At) - u(x, t)]

where c is the heat capacity and p is the density ( p r unit length) of the material from which the rod is made. (Thus, ps Ax is the mass of the element AB.)

A comparison of (20.2) and (20.3) shows that I

au(x, t + 02 At) = K a%(x + O1 AX, t) CP at ax2 9

and if we pass to the limit as At + 0, Ax + 0, then

where a2 = K/cp. In this way, we obtain the equation for heat flow (or heat conduction) in a rod.

We now pose a variety of problems, corresponding to different conditions imposed on the ends of the rod.

21. Heat Flow in a Rod With Ends Held at Zero Temperature

This problem consists in finding the solution of equation (10.4), with the boundary conditions

(the ends of the rod are at x = 0 and x = I), and with the initial condition

where f(x) is a given function. Equation (20.4) is a special case of equation (1.2) of Part I, and hence all the considerations given there apply to the present problem.

Thus, we look for particular solutions of the form

Page 312: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

298 THE m ~ r n o ~ METHOD AND rrs APPLICATIONS CHAP. 9

which do not vanish identically and which satisfy the boundary conditions 2 . 1) Substituting this expression in (20.4) gives

@T' = aWnT,

Therefore, we have

db" + ?,a = 0, T' + a2PT = 0.

The solution of (21.3) is

@(x) = Cl cos Ax + C2 sin Ax,

and because of the condition (21.1), we must require that

O(0) = Cl = 0,

4(l) = C2 sin M = 0.

Therefore, assuming that C2 # Oj we obtain M = ra, where n is an integer. If we set Cz = 1, then

*

?cn An = TY

ZIMC @(X) = sin A,,x = sin (n = 1,2,. . .).

For X = )k, equation (21.4) gives .

Hence, the functions

a%%9 nnx --r u~(x, t) = An sin 7 e 0 (n = 1,2,. . .) represent particular solutions of (20.4) which satisfy the boundary conditions.

To satisfy the initial condition, we form the series

00 A%= 7W1X - - t u(x, t) = 2 An sin 7 e f l

n= 1

18 We leave it to the reader to decide why the constant is chosen to be negative here (eee Sec. 12).

Page 313: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 21 THE EIGENFUNCTION METHOD AND ITS APPLICATIONS 299

and require that 00 m x

U(X, 0) = 2 An sin - = f (x). n= l 1

Thus, we have to expand f(x) with respect to the system {sin ( x I ~ x / ~ ) ] . A calculation of the Fourier coefficients gives

i.e., the solution of our problem is given by the series (21.9, where the coefficients A, are determined from the formulas (21.6). Because of the presence of the factors

it is easy to see that the series (21.9 is uniformly convergent for t 2 to > 0, for any to > 0. The same is true for the series obtained by term by term differentiation of (21.5) with respect to x and t (any number of times). Therefore, the sum of the series (21.5) is continuous, and the term by term differentiation is legitimate (cf. Sec. 10).

22. Heat Flow in a Rod with Ends Held at Constant Temperatures

This problem consists in finding the solution of equation (20.4), with the boundary conditions19

and with the initial condition

24x9 0) = f (XI.

We look for a solution in the form of a series QO

mix u(x, I ) = 2 T') sin -9

n= 1 1

where

2 1 m x TM(t) = 7 Jo U(X, t) sin - I dx.

19 The boundary conditions in this problem, and also in the problem of Sec. 23, have a different form from those considered previously. We show below how to deal with cases like these.

Page 314: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

300 THE EIGENFUNCTION METHOD AND ITS APPLICATIONS CHAP. 9

Integrating by parts twice, we obtain

12 la% m x --I- xzn* 0 8x2 sin - dx. I

Since u(x, t) satisfies equation (20.4) and the conditions (22.l), we have

1 1 1 mx -Tn = - 2 1' 2 sin 7 dx. nn [A - (- l)nBl - =%zn2 at

Differentiating (22.4) with respect to t, we obtain

2 1 au nnx Ti = pin--dx, 1

so that

This equation has the solution

To satisfy the initial condition (22.2), we require that

a nnx U(X, 0) = 2 Tn(0) sin = f (x). n = 1

A calculation of the Fourier coefficients of f(x) with respect to the system {sin ( m c / l ) } gives

A - (-1)"B = - / 2 1 mx Tn(0) = A. + 2 f(x) sin - dx, xn 1 6 0 I

and hence

2 i xnx A, = 7 j" f(x) sin - dx - 2 A - (- 1)"B I

. xn (22.7)

Thus, the solution of our problem is given by the series (22.3), where the Tn are determined from the formulas (22.6) and (22.7).

Page 315: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 23 THE EIGENFUNCTION METHOD AND ITS APPLICATIONS 30 1

23. Heat Flow in a Rod Whose Ends are at Specified Variable Temperatures

In this problem, it is required to find the solution of equation (20.4), with the boundary conditions

~ ( 0 , 1) = MI, ~(1st) = 44) (23.1)

(where cp and + are given), and with the initial condition

U(X, 0) = f (x). (23.2)

We again look for a solution in the form of a series (22.3). Repeating the argument of Sec. 22, we find for Tn(t) the equation

which is the same as formula (22.5), except that A and B have been replaced by cp and 9. Solving this equation, we obtain

To satisfy the condition (23.2), we require that a0 znx

U(X, 0) = 2 Tn(0) sin 7 = f (x). n o l

A calculation of the Fourier coefficients of f(x) with respect to the system sin {(xnxll)} gives

2 1 m x T'(0) = A, 7 I. f(~) sin 7 d ~ .

Therefore, the solution of our problem is given by the series (22.3), where T, is determined from the formulas (23.3) and (23.4).

24. Heat Flow in a Rod Whose Ends Exchange Heat Freely with the Surrounding Medium

If the surface of a body exchanges heat with a surrounding gaseous medium, then the amount of heat flowing through an area s in the time At is given by the formula

Q = H(u - uo)s At, (24.1)

Page 316: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

302 THE EIGENPUN~ON WTHOD AND m APPLICATIONS CHAP. 9

where u is the temperature of the body, uo is the temperature of the sur- rounding medium and His a constant called the em~sivity.

In the case of heat flow in a rod whose lateral surface is insulated, but whose ends exchange heat freely with the surrounding medium, a comparison of (20.1) and (24.1) leads to the boundary conditions

for x = 0, and

for x = I. Setting h = H/K (h > 0), these conditions become

First, let us assume that uo = 0. The boundary conditions then take the form

[s - hu] = 0, ax x-0

[g + hu] = 0, -1

while the initial condition is

as before. Following our usual method, we look for particular solutions of equation (20.4) of the form

satisfying the conditions (24.3). Substituting this expression in (20.4) gives

whence

0" T' = = Q, a2T - A = const,

so that

Page 317: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 24 THE EIGENmJNCTION METHOD AND ITS APPLICATIONS 303

To satisfy the boundary conditions (24.3), we must obviously require that

Thus, we have arrived at a boundary value problem for equation (24.6). with the boundary conditions (24.7). According to Sec. 5, all the eigen- values of this boundary value problem are positive. Therefore, we can write A2 instead of X; thus, instead of the equations (24.6). we obtain

The solution of (24.8) is

O(x) = C, cos Ax + C2 sin Ax,

and by (24.7). we must have

A(- Cl sin M + C2 cos M) + h(C1 cos M + C2 sin M) = 0.

Therefore

so that

2M tan M = Xz - h2'

The positive roots of this equation give the eigenvalues of our problem. It is worth pointing out that these roots are the abscissas of the points of inter- section (in the @-plane) of the function p = 2/tan M and the hyperbola p ' (A2 - h2)pdi.

Now, let & be the nth positive root of equation (24.11). According to (24. lo), we can set Cl = &, C2 = h. Then, the eigenfunctions become

For A = A , the solution of equation (24.9) is

This leads to the following particular solutions of (20.4):

u~(x, t) = A,(& cos )Wr + h sin h , ~ ) ~ % t (n = 1, 2, . . .).

Page 318: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

CHAP. 9

To satisfy the initial condition (24.4), we form the series

U(X, t) = An& cos )Uc + h sin AS)-%, (24.12) n = l

and require that QO

U(X, 0) = 2 Ah& a s &x + h sin h s ) = f(x). n= 1

A calculation of the Fourier coefficients of the function f (x) with respect to the system20

{a) = {A,, cos ?g + h sin Sr)

gives

Thus, the solution of our problem is given by the series (24.12), where the coefficients are determined from (24.13).

The integral in the denominator of (24.1 3) can be calculated as follows: The equation

implies that

h# = -@,a;. Therefore, we have

But

C=h,+osk,,x+ hsinAp, 0; = -h: sin A,,x + M, cos ?t,,x,

which means that

+ = A: + h2h2,

and hence

According to Sec. 4, this system is orthogonal.

Page 319: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 24 THE EIGENFUNCTION METHOD AND ITS APPLICATIONS 305

This, together with (24.14), implies

On the other hand, it follows from the boundary conditions and (24.15) that

at both x = 0 and x = 1. Therefore, writing the boundary conditions in the form

[On@; - h@:L0 = 0, 4 + = 0,

and using (24.17), we find that

[anO;X:; = -2h1:.

Substituting this expression in (24.16) gives

Thus, instead of (24.13), we can write

2 Cf(x)(ln cos + h sin 1 s ) dx

Next, we consider the case wh&e the end of the rod at x = 0 exchanges heat with a medium at the temperature s, while the end of the rod at x = 1 exchanges heat with a medium at the temperature ul. This problem can be reduced to the problem just solved by making the substitution

where the function v = v(x) depends only on x and satisfies the equation

v" = 0, (24.19)

with the boundary conditions

[v' - h(v - uU~)]-~ = 0, [v' + h@ - ul)Jml = 0, (24.20)

while w satisfies the equation

Page 320: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

CHAP. 9

with the boundary conditions

[g - hw] = 0, -0

[$ + hv] = 0 -1

and the initial condition

According to (24.19). the function v = v(x) has the fonn

where the constants A and B are determined from the conditions (24.20). This leads to the system of equations

which can be solved very easily. Then, the boundary value problem for w is of the type discussed above.

25. Heat Flow in an Infinite Rod

In the case of an infinite rod, there are no boundary conditions, and the problem reduces to finding a solution of equation (20.4) which is defined for all x and t > 0, and satisfies the initial condition

As usual, we look for particular solutions of the form

u(x, t) = @(x)T(t).

Substituting this expression in (20.4) gives

OT' = aWUT,

whence

so that

Page 321: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

These equations have the solutions

@(x) = Cl cos X* + C2 sin hx, T(t) = C3eQet.

Therefore, the required particular solutions can be written in the form21

t((X, t; A) = (A cos Ax + B sin A x ) d e t .

The constants A and B, being arbitrary, can be regarded as values of functions A = A(1) and B = Be). It will be recalled that for discrete values of A (as in the case of a finite rod), we formed an infinite series of particular solutions, and we then chose the coefficients of the series in such a way as to obtain a solution satisfjhg the initial condition. In the present case, A varies continuously, and we set

(A@) cos hx + B e ) sin k c ) d e t A. (25.9

If we can differentiate this function behind the integral sign (once with respect to t and twice with respect to x), then the function u(x, t) wilfbe a solution of equation (20.4). In fact, we then have

To satisfy the initial condition, we require that

U(X, 0) = (AN cos h~ + B(X) sin h) A = =(XI. 0

This will l the case if we require that f(x) can be represented as a Fourier integral (see Ch. 7, Sec. 5), a sufficient condition for which is that f(x) be a piecewise smooth and absolutely integrable function on the whole x-axis. With these assumptions

A@) = '5" x -00

f(v) cos nt) do,

1 0 B(V = ; j ' ~ ( o ) sin L dv

[see equations (5.5) to (5.7) of Ch. 71. - -

21 If we chose the constant to be positive in (25.2), then the exponential factor which falls off as t increases would be replaced by an exponential factor which increases without limit as t increases, contrary to the physical meaning of the problem.

Page 322: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

308 THE EIGENFUNCTION METHOD AND ITS APPLICATIONS CHAP. 9

For such A@) and B(A), the integral (25.5) can be differentiated behind the integral sign any number of times with respect to x and t. In fact, because of the presence of the factor ea2t in the integrand of (25.5) and because of the inequalities

the integral (25.5) and the integrals obtained from it by differentiating the integrand any number of times with respect to x and t are uaiformly convergent for t 3 to > 0 (where to > 0 is arbitrary). This follows from the inequalities

I (A@) cos hx + B(A) sin la) Ie-A2t $ 2Ce-3Q2t 6 2Ce-47QQ0,

[(Am cos hx + B@) sin hx)e-a%@

< 2Chne-aQQ ( 2(3ne-Q2t09 \

[(A@) cos ?a + B ) sin hx)e-aQet] I < 2CabAb e-a2t < 2Cah~%-~A~to,

and the fact that the majorizing functions on the right are integrable in A from 0 to m. Thus, we need only apply Theorems 4 and 3 of Ch. 7, Sec. 6. It should be noted that although our argument shows that u(x, t) is a solution of equation (20.4), it does not show that

lim u(x9 t) = f(x). t+o

However9 this relation is true and can easily be proved. Using (25.6), we can rewrite the solution of our problem as

To further transform this expression, we begin by proving that it is possible to change the order of integration. To show this, we note first that for every E > O

for sufficiently large I (where t > 0 is fixed). Therefore

Page 323: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 25 THE EI~ENPUNCTION METHOD AND ITS APPLICATIONS 309

from which it follows that the integral in the left-hand side converges to zero as 1 -+ oo. But then

do r f(v) cos h(x - v ) e - A Q h 7t -a0 0

1 - lim - 100 d~ J1 f(v) cos A(X - - - ) e a t A ba, 7c -a, 0

1 1 = lim - j' I" f(u) cos x x - o)e-a2t (iD = U(X, t ) (+a0 7t 0 -a,

[see (25.91. Here the change in order of integration is legitimate, because the integral

is uniformly convergent in h for 0 ( A ( I. This follows from the fact that the integrand is dominated by (f(v)l (see Theorems 4 and 2 of Ch. 7, Sec. 6).

Thus, we can write

instead of (25.7). It turns out that the inner integral can be evaluated. In fact, set

akd i = z, 7(x - v) = pz

so that

dz X - l , dx=cn/i. p = -0

d t

Then we have

1 e-fi cos pz dz = - I(p). (25.9)

d t

Differentiation with respect to p behind the integral sign gives

~ ' ( p ) = - Joa, e-z2 z sin pz dz;

this differentiation is legitimate because the resulting integral is uniformly convergent in p.

We now integrate by parts, obtaining

1 "" - 5 lom ' 2 2 QH p dz c ~ ( p ) = 2 [e-z2 sin pzlZ-,, - f I(P).

Page 324: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

3 1 0 THE EIGENFUNCIlON METHOD AND ITS APPLICATIONS CHAP. 9

It follows that

I(p) = ce-w2f4.

To find C, we set p = 0. This gives

an integral which, as we know, equals */;; (see Ch. 8, Sec. 8). Therefore

I(p) = *& e - ~ ~ 1 4 ,

and by (25.9)

Substituting this expression in (218), we finally find

1 00 (x- v)= f ( o l e - z dv.

On the one hand, formula (25.10) shows that as t increases, u(x, t) + 0, i.e., the heat "spreads out" along the rod. On the other hand, (25.10) shows that the heat is "transmitted" instantaneously along the rod. In fact, let the initial temperature be positive for xo < v < xl and zero outside this interval. Them the subsequent distribution of temperature is given by the formula

from which it is clear that u(x, t) > 0 for arbitrarily small t > 0 and arbi- trarily large x.

26. Heat Flow in a Circular Cylinder Whose Surface is Insulated

Let the axis of a circular cylinder of radius I be directed along the z-axis, and let its ends be insulated (or let the length of the cylinder be infinite). Suppose that the initial temperature distribution and the boundary conditions are independent of z. Then, it can be shown that the equation for heat flow is

where a2 = K'cp, K is the conductivity of the material from which the rod is made, c is its heat capacity, and p its density. Thus, the temperature is

Page 325: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

se~. 26 THE EIGENFUNCIlON METHOD AND ITS APPLICATIONS 3 1 1

independent of z (a fact which is of course a consequence of the assumptions just made), and we are essentially dealing with a problem in the plane. If we go over to polar coordinates by setting x = r cos 0, y = r sin 8, then, instead of equation (26.1). we obtain

We now assume further that the initial and boundary conditions are independent of 0. Then, obviously u is a function only of r and t, and the heat flow equation takes the form

We also assume that the surface of the cylinder is insulated from the sur- rounding medium, i.e.,

(absence of heat flow), and that the initial temperature distribution is given by the condition

We look for particular solutions of the form

u(r, t ) = R(r)T(t).

Substituting this expression in (16.2) gives

whence

so that

Equation (26.5) is the parametric form of Bessel's equation, with index p = 0 (see Ch. 8, Sec. 11). Its general solution is

Page 326: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

3 12 THE BIOENPUNCT~ON METHOD AND ITS APPLICATIONS CHAP. 9

S i n a Yd?w)-+ a, as r+ 0, we have to set C2 = 0. Taking Cl = I , we find from the boundary condition (26.3) that

J W ) = 0.

Therefore, p = Al is the nth positive root of the equation J&((IL) = 0. We write

CiLn &=7

where p,, = y is the nth positive zero of the function Jh(). Setting X = A,, in equation (26.6), we find

Tn(t) = ~,,e-a%t (n = l,2, . . .). (26.7)

Thus, for equation (26.2), subject to the condition (26.3), we have found particular solutions of the form

un(r, t) = A,J&)pWt (n = 1,2, . . .). (26.8)

We now form the series

and to satisfy the initial condition (26.4), we require that

A calculation of the Fourier coefficients off(r) with respect to the system ~JdLr)} Bives

(see Ch. 8, Sec. 24). Therefore, the solution of our problem is given by the series (26.9), where the coefficients A, are determined from the formula (26.1 1).

27. Heat Flow in a Circular Cylinder Whose Surface Exchanges Heat with the Surrounding Medium

This problem reduces to solving equation (26.2) with the boundary condition

Page 327: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

SEC. 27 THE BIGENFUNCTION METHOD AND 1TS APPLICATIONS 3 1 3

and with the previous initial condition

Repeating the argument of Sec. 26, we again obtain equations (26.5) and (26.6), and we again find that

R(r) = Jo&)*

The condition (27.1) gives

AJh(M) + h J&I) = 0

Therefore, the number p = M must be a root of the equation

We now write

wherep,, is the nth positive root of equation (27.3). For A = )k (n = 1,2,. . .), the solution of equation (26.6) is given by (26.7), and (26.8) gives particular solutions of (26.2) satisfying the condition (27.1). We again form the series (26.9); and require that the relation (26.10) be satisfied. A calculation of the Fourier coefficients of f(r) with respect to the system {Jo(A,,r)} leads to the formula

(see Ch. 8, Sec. 24). Thus, the solution of equation (26.2), subject to the conditions (27.1) and (27.2), is given by the series (26.9), where the coefficients are determined from (27.4), and the numbers pn are the roots of equation (27.3).

28. Steady-State Heat Flow in a Circular Cylinder

We now assume that a constant temperature is maintained on the surface of the cylinder, and that the distribution of temperature is independent of z.

Page 328: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

3 14 THE EIOBNPUN~~ION METHOD AND ITS APPLICATIONS CHAP. 9

Then after a suaadently long interval of time, a definite temperature is established at every point of the cylinder, i.e., the function u ceases to depend on t. Thus, instead of equation (26.1), we have

or in polar coordinates

Let the temperature on the boundary be specified by the condition

and look for particular solutions of the form

Substituting this expression in (28.1) gives

whence

so that

The solution of (28.5) is

@(B) = A cos U) + B sin A0.

It follows from the physical meaning of the problem that the function @(0) must have period 27t, and hence X must be an integer. (Incidentally, we note that @(0) would not have been periodic if we had taken the constant h (28.3) to be positive.) Thus, we write

For h = n, equation (28.4) takes the form

which is a second order linear differential equation. It can be verified by direct substitution that the functions F and F satisfy this equation. There- fore, for n > 0, the general solution of (28.7) is

R,, = C#" + D,p. (28.8)

Page 329: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Since rn -+ a, as r -+ 0, we have to set Dn = 0. For n = 0, we easily find that

Ro = Co + Do In r, (28.9)

and hence we must again take Do = 0. Using (28.6), (28.8), (28.9), and the conditions D, = 0 (n = 0,1,2, . . .), we can write the particular solutions of (28.1) as

Un(r, 0) = (an COS tZO + pn sin tZO)PJ (n = l,2, . . .),

We now form the series

~ ( r , 0) = + 2 (gl cos & + pn sin n0)~ , n=l

and to satisfy the boundary condition (28.2), we require that

u(1,0) = 4 + 2 (a,, cos n0 + p, sin n8)ln = f (0). 2 n = l

A calculation of the Fourier coefficients off (0) gives

1 pJ"=- f(0)sin&&=bn XJ,

(n = 1,2,. . .), so that

Therefore 00

~ ( r , 0) = 7 + 2 (a,, MS 120 + bn sin dl) (5r. (28.10) n= 1

For r < I, this series can be differentiated term by term any number of times with respect to r and 0, since each resulting series is uniformly convergent for 0 ( r 6 rh where ro < 1 is arbitrary. It follows that (28.10) actually gives the solution of equation (28.1).

This solution can be given a more compact form if we use the Poisson integral (see Ch. 6, Sec. 7). Then we have

1 = 12 - rz u(r, 0) = - 2rr J-.f(') P - z r cos (t - 01 + P dtm

Page 330: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

CHAP. 9

Moreover

wherever f(0) is continuous, i.e., the solution just written satisfies the boundary condition (28.2).

PROBLEMS

1. Find the eigenvalues and normalized eigenf'ctions of the differential equation @"(x) + X@(X) = 0 on the interval [O, 11, subject to the following boundary conditions:

a) @(O) = @(I) = 0; b) @(O) = @'(I) = 0; c) @(O) = @(I) + hO'(1) = 0.

2. Consider a string of length 1, tension T and linear density p, fastened at the points x = 0 and x = I. Suppose that at time t = 0, the point x = c (0 < c < I) is displaced by an amount h and then released (the "plucked string"). Write the initial conditions and find the subsequent motion of the string.

3. Consider a string of length 21, tension T and linear density p, fastened at the points x = f 1. Let the initial position of the string be a parabola which is symmetric with respect to the center of the string, with maximum initial dis- placement equal to h, and let the initial velocity of the string be zero. Write the initial conditioh and find the subsequent motion of the string.

4. Consider a string of length 21, tension T and linear density p, fastened at the points x = f I. Suppose that at time t = 0, the string receives an impulse of magnitude P at its midpoint. Find the subsequent motion of the string.

Hint. Solve the problem with the initial conditions u(x, 0) = 0,

P for 1x1 B 8,

at for e < 1x1 g I, and then pass to the limit s -+ 0.

5. Consider a string of length I, tension T and density p, fastened at the points x = 0 and x = I. Let the string be initially at rest in its equilibrium position. Suppose the section of the string between xl and xz (0 d XI < xz < /) is acted upon by a periodic perturbing force F(x, t) = pA sin at (see Sec. 13). Find the subsequent motion of the string. Verify that when xl = 0, xz = 1, the answer reduces to formula (1 3.8).

6. Let a amcentrated periodic perturbing force P = A sin at act upon the point x = c of the string of Prob. 5. Find the subsequent motion of the string.

Hint. Pass to the l i t XI-+ c, xz -+ c in Rob. 5.

Page 331: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

PROBLEMS THE EIGENFUNCTION METHOD AND APPLICATIONS 3 17

7. Consider a rod of length I, Young's modulus E, cross section s and density p, fmtened at the end x = 0. Suppose that the rod is stretched by a f m F acting on the end x = I, and then is suddenly released at time t = 0. Write the initial conditions and find the subsequent longitudinal vibrations of the rod.

Hht. The amount by which the rod is initially stretched is I;l/B.

8, Let the free end of the rod of Prob. 7 receive a sudden impulse P at time t = 0. Find the subsequent longitudinal vibrations of the rod.

Hht. Solve the problem with the initial conditions

for 0 < x < I - e, au(x, 0) =

at for l - e < x b I,

and then pass to the limit e + 0.

9. Suppose the f m end of the rod of Prob. 7 is acted upon by a periodic per- turbii force A. sin a t . Find the subsequent longitudinal vibrations of the rod, assuming that the rod is initially at rest.

10. Find the linear combinations of the modes u13 and usl of a rectangulas membrane corresponding to the nodal lines of Fig. 51 (c) of Sec. 17. Write the equation of the "nodal ring" indicated by the fourth sketch in Fig. 51(c). Is this curve actually a circle?

11. Consider a circular membrane of radius I, tension T and density p per unit area. Find the radial vibrations of the membrane if it receives a'sudden impulse Pat tim t = 0 distributed over the circle r G e, assuming that the membrane is originally at rest in its equilibrium position.

Hint. Solve the problem with the initial conditions

P for 0 4 r G s, at for e < r < 1.

12. Find the radial vibrations of the membrane of Prob. 11 if it is acted upon by a periodic perturbing force F = A sin a t per unit area, uniformly distributed over the entire membrane, assuming that at time t - 0 the membrane is at rest in its equilibrium position.

13. Consider an infinite slab of width 21 bounded by the planes x = f I, made of material with thermal conductivity K, spec& heat c and density p. The slab is first heated to temperature To and then at time t = 0, its faces are held at zero temperature. Find the subsequent temperature distribution in the slab.

14. Let the slab of Prob. 13 have an initial temperature distribution u(x, 0) = f(x). Find the subsequent temperature distribution if starting from time t = 0, the slab exchanges heat h l y with the surrounding medium which is at zero temperature. (Let the emissivity of the slab be H (see Sec. 24.))

Page 332: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

8 THE EIOENFUNCTION METHOD AND ITS APPLICATIONS CHAP. 9

15. Fmd the steady-state temperature distribution in an infinite rectangular prism bounded by the planes x = 0, x = a, y = 0, y = b, if the forces formed by the planes x = 0, x = a, y = 0 are at zero temperature, while the face forrned by the plane y = b has the temperature distribution u(x, b) = f(x), where 0 6 x < a. (Draw a f i m a ) Specialize the answer to the case f(x) = To.

Page 333: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

ANSWERS TO PROBLEMS

2 00 nsinnx C) ~hino~= 2 (-1PbT2 (-x < x < z); x

n=1

2 b) sinh ax = - sinh ax

n ( -SC< x < SC).

K

00

4. a) sin rur = cos2nx l -wsax[t Ts

+ & n=l 2 a2-4n2]

Page 334: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

320 ANSWERS TO P R O ~ ~ M S

(What happens when a is an integer?)

b) AX) = - (0 6 % 6 z), n= 1

except for the value x = h, where the sum equals 112. (Why?)

C) f(x) = 2[;+ 'It 2 r ) 1 a s a ] ( 0 4 x 4 4 . n=1

t zx 4 n r a ) f(xj=-sin--- 5 ( - 1 ~ ~ ~ ~ 2 2- (0 6 x < 1), sin -

% n - l - 1 1

except for the value x = 112, where the sum equals 112;

b) Ax) = sin - n=2 n2 - 1 I

except for the value x = 112, where the sum equals 0.

Hence

I2={f(X) - f (; + x)} an = - 2% 0

so that

PO 00 sin nx sin nx b) x3=2x2 Z ( - l p + l - + l ~ 2 (-l)n-. n= 1

n n= 1 n3

Page 335: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

ANSWERS TO P R O B L ~ S 32 1

CHAPTER 2

3. By the Cauchy inequality

Now integrate. The sharper estimate is obtained by direct use of the Schwarz inequality.

4. Use the Schwarz inequality with one function equal to I I + einrx + + elnkxl

and the other equal to 1.

5. Use the Schwarz inequality.

7. Suppose g(x) is orthogonal to all the cpl(x), and let P,(x) = ancpo(x) + + anq,,(x) be such that

Then Pg 112 + 11 Pn 02 < 1 In, so that llg 112 = 0, i.e., g(x) = 0, since g(x) is continuous.

8. a) No orthogonal function exists, since if g were such a function and if

then 0 = co + cl = co + c2 = co + c3 , i.., - co = cl = c2 = 9

which violates Bessel's inequality (Sec. 6) unless all the ci are zero; b) No orthogonal function exists, by a similar proof; c) An orthogonal function g exists. In fact, consider the continuous function

9. a) If g = aocpo + + an% is zero, then

i.e., a. = al = = an = 0 [for the definition of (9, (k), see Prob. 10 and Ch. 2, Sec. 101; b) A polynomial of degree n has at most n real zeros, unless it vanishes identically.

11. Use repeated integration by parts.

Page 336: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

CHAPTER 3

2. a) Neither limit exists; b) f:(O) = 0, f'(O+) does not exbt;

c) C ( 0 ) = f ( O + ) = 0.

3. This follows from (6.3) just as in the proof of Sec. 6.

Moreover, we haw

then

by the Cauchy inequality (see Ch. 2, Prob. 2).

6. We have

It follows from (2.9) that

b - a j) + a ~ n r c o s 2 0 ~ - sinnwsin20~)d~ > - 2

if n 3 N. Therefore

b - a < M(6 - a), that 2 pa < 4M. n=N

Page 337: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

7. a) The formula

follows from equations (3.2) and (4.1) after some manipulation ; b) If

1 sin nx D'.(x) = D,(x) - 2 cos nx = 2 tan (~12)'

then

as n + oo (why ?). Moreover

1 ~:(t)) dt = r 0 (I t - I mt i) sin nt dt+ o

asn+oo. (Why?)

Thus

sin nt ~ n ( t ) dt = I 0 7 dt + un(x),

where u,,(x) -+ 0 as n -+ oo . It

C) 1; dt = follows from b) when n + oo.

8. Use preceding problem. The inequality can be derived by inspecting the area under the curve y = sin tit.

CHAPTER 4

1. a) x # 2h; b) for all x; c) x # 2kx.

2. For x # 2kz. No*

3. a) x # (2k + 1)x; b) x # 2kx; c) for all x; d) for all x.

4. a), b) for x # (2k + 1)x; c) for x # 2kls/3; d) for x # (2k + I)lc/2.

6. a) sin (cos x) cosh (sin x); b) cos (cos x) sinh (sin x).

7. a) cos (cos x) cosh (sin x); b) sin (cos x) sinh (sin x).

8. a) (1 + cos x) ln 2 cos - + sin x; ( 3 ; X

b) 2 (1 + cos x) - sin x In ( 2 cos ) - ( - n < x < x ) .

Page 338: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

b) sinxlog 2cos- --sinx ( - r < x < x ) . 3 : 1 1O.a) cosxln

[ ( 3 + , 2 1 ~ ] + r i n p x [ Z + X - $+I; 11. a) cos px In 2 sin - n u 1

(0 < x < 2%).

12. Use Abel's lemma (Sec. 1).

13. Imitate the proof of Abel's lemma.

14. Assume that 0 < xo < r. Sice cos2 0 + 'lccw 01, we have

But 2 cos2 roxo = 1 + cos kO, and by Theorem 1 of Sec. 3,

Therefore, the series

CHAPTER 5

Page 339: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

3. By making the odd extension of f(x) onto the interval [-IC, 01, we obtain an odd function F(x) with equal (zero) values at the end points of the interval [- IC, n]. Moreover, F'(x) is obviously an even differentiable function, P(x) is an odd f m - tion which again has equal (zero) values at the end points of the interval [-IC, n], and F8"(x) is a continuous even function. Therefore, F(x) and its k t three derivatives can be extended continuously over the whole x-axis, with P(x) a smooth function But we know that the Fourier coefficients of a continuous function and those of its derivative are related by the formulas

[see (8.3)]. Therefore, in our case

Moreover, for a smooth function, the sum of the squares of the absolute values of the Fourier coefficients converges (Ch. 3, Sec. 10). Since

it follows that the last series converges. This in turn implies the uniform conver- gence of the &es obtained by one or two term by term differentiations of the original series, and the uniform convergence of these series guarantees that the term by term differentiation is legitimate.

1 a0 cos ax 6. a) f'(x) = - 2 - 2 n Z + n + 1. n= 1

1 1 00

b) f(x) = - - *- - cos nx

n=l

Page 340: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

7. F(x) = cl cos x + c2 sin x + + (0 < x < =),

where

1 1 + - = 2, since 1

n=2 n2-1 2

8. a) (In each case f(x) denotes the sum of the series.)

aD 7 c - X sin nx

=-' 2 Cm (0 < X < 27C) n= 1

[see formula (2) of Sec. 12.1

7 c - x sin nx d@ + 1)

(0 < x < IT).

x (- l)rc+l sin nx

n- 1 n2(n + a) (-lc < x < x);

Page 341: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

9. Use Theorem 1 of Sec. 3.

10. Use Theorem 1 of Sec. 8 and equation (3.1).

11. To derive formula (I), use formula (3.1). Then, substitute h = =/2N in f m u l a 0. 12. a) Set N = 2k and apply the last part of the preceding problem, keeping only

terms with index n 3 N/2. (For such terms, sin2 (m12N) 3 +.)

b) Use the Cauchy inequality (Ch. 2, Rob. 2);

c) Sum the inequalities b) with respect to k.

CHAPTER 6

4. a) a = +; b) a = 1 (1 + pI2

9 the series converges.

7. a) 3; b) +; c) +; d) Not summable by arithmetic means, but Abel-summable to 0; e) Not summable by arithmetic means, but Abel-summable to +.

8. Use the relation s, = (n + l ) ~ , + ~ - na,.

9. a) Use the relation tR = (n + l)s, - (n + lb,,+l; b) This follows from a) and the theorem of Sec. 2.

CHAPTER 7

sin rrh . @(k) = - - A2'

- sinh- X m s h c) @(A) = d 2 / x 2h A2

; d) @(A) = d G (1 + 1 9 2

2 1 - COS'ZCX 6. a) f(x) = - 2 1

X x ; b) f ( ~ ) = , ~ +x2;

2 2ax c) f(x) = ; ,, + ,,,*

7. Imitate proofs for the discrete case (Ch. 3, Prob, 4).

Page 342: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

forx < 0, 8. b) Mx) = ce-x 0;

(0 for x < 0, +x3 for0 < x 4 1, -Hx3 - &C + 4) for 1 g x < 2,

10 for x 2 2.

for x < 0, f(x) = &e-X fozx s 0;

CHAPTER 8

1. Setting p = 2, we reduce the equation to the fonn (2.2). Then, the substiltion u = x2y [see (241 reduces the equation to the following Bessel's equation:

It follows at once that

- - (Zn - I)(h - 3). . * 5 * 3 e 1 cX 2"

(see Sec. 3).

where B = const, = const and p(x) is bounded as x -+ 00.

4. J;(x) = A sin (x + o) <x)

fi + x where A and o are the same constants as in the asymptotic expression for Jp(x), and ~(x) is bounded as x+ ao.

6. Use formula (7.1).

9. Use the precedii problem and formula (2.9) of Ch. 3.

10. We have

lo1 x-p+lJJ~g)& = - ,$p+2 j: t-H1JP(t) dt.

Page 343: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Replacing p by p - 1 in formula (7.2), we obtain d - [t-P+lJp-l(t)] = - t-p+lJP(t). dt

Therefore

[see (4.3)], and hence

The series converges to XP if p 3 - +, p > H.

for 0 6 x < 2.

CHAPTER 9

In Problems 2 to 6, a2 = Tip.

2h 12 nxx nxat 2. sin - I a s -* I

Y ( x y t ) = ~ ~ ( ~ ~ - m s - I mX2) I sin nxx T * w sin ant - w, sin a t y

5C n = l n a n ( d - a

where on = mall.

2A " nxc nxx w sin ant - on sin o t a t ) = - x s i n ~ sin -* I

on(02 - 08) n - 1

In Problems 7 to 9, a* = E/Q.

Page 344: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

10, ul3, u31, ~ 1 3 - US,, u13 + ~ 3 1 . The nodal ring has the equation cos27cx + cos2Ity + 1 r 0,

and is not quite circular.

2P 11 U(r , t )= - a J l O . n J 0 ( r ) I+ = @ ~ c z ~ d m

Jo p,, 7 sin

where p,, is the nth positive root of the equation Sob) = 0 and c2 = T/p.

where (An is the nth positive root of the equation Jdp) = 0 and o, = &I.

Kt where s = -. CP

exp { - (a,%/i 2)) 11 m) cos & 14 +x, = f 2 1 + (sin 2at,IhJ n o 1 -1

l a " (BslJ) exp ( - ( B ~ J P ) } $ fg) sin in &, + 7 n o t 2 1 - sin@^?^^) -I

where s = (Ktlcp), a, is the nth positive root of the equation tan x = ZHIx, and pn is the nth positive root of the equation tan x = -x/lH.

" sinh (myla) 15* u < s y ) = z ~ sin r f(6) sin d - 4, a ~ = ~ sinh (msbla) a 0 a

where sinh x is the hyperboc sine (cf. Rob. 3 of Ch. 1). In the special case

4To 2 U(x, Y ) = - siah [(2n + 1) =y]a] sin [(2n + I ) =lo). It ~ E O sinh [(2n + 1 ) &/a] h+1

Page 345: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

BIBLIOGRAPHY

Bochner, S., Lectures on Fourier Integrals, translated by M. Tenenbaum and H. Pollard, Princeton University Press, Princeton (1959).

Bochner, S. and K. Chandrasekharan, Fourier Transforms, Princeton University Press, Prrrnceton (1949).

Byerly, W. A., An Elementary Treatise on Fourier's Series and Spherical, Cylindrical, and Ellipsoidal Harmonics, Ginn and Co., Boston (1893).

Carslaw, H. S., Introduction to the Theory of Fourier's Series and Integral, Macmillan $ Co., Ltd., London (1930), reprinted by Dover Publications, Tnc., New York.

Carslaw, H. S. and J. C. Jaeger, Conduction of Heat in Solih, second edition, Oxford University Press, New York (1959).

Churchill, R. V., Fourier Series and Boundary Value Problems, McGraw-Hill Book Co., Inc., New York (1941).

Franklin, P., Fourier Methods, McGraw-Hill Book Co., Inc., New York (1949), reprinted by Dover Publications, Inc., New York.

Goldberg, R. R., Fourier 7Fansforms, Cambridge University Press, New York (1961). Jackson, D., Fourier Series and Orthogonal Polynomials, Carus Mathematical

Monograph No. 6, Math. Assoc. America (1941). Jeffrey, R. L., Trigonometric Series, Univerbity of Toronto Press, Toronto (1956). McLachlan, N. W., Bessel Fullctions For Engineers, second edition, Oxford Univer-

sity Press, New York (1955). Rogosinski, W. W., Fourier Series, translated by H. Cohn and F. Steinhardt, second

edition, Chelsea Publishing Co., New York (1959). S d o n , I. A., Fourier Transforms, McGraw-Hill Book Co., Inc., New York (1951). Titchmarsh, E. C., Introduction to the Theory of Fourier Integrals, second edition,

Oxford University Press, New York (1948). Titchmarsh, E. C., Eigenfunction Expansions Associated with Second-Order Diger-

ential Quations, Oxford University Press, New York, Part I (1946), Part I1 (1958). Watson, G, N., A Treatise on the Theory of Bessel Functions, second edition,

Cambridge University Press, New York (1945). Wiener, N., The Fourier Integraland Certain of Its Applications, Cambridge Univer-

sity Press, New York (1933), reprinted by Dover Publications, Inc., New York. Zygrnund, A., Trigonometric Series, in two volumes, Cambridge University Press,

New York (1959).

Page 346: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

INDEX

Abel's h a m , 97 Cauchy inequality, 63 Abel's method of summation, 162 Cauchy principal val\~e, 11f9 Absolutely i n w b l e W o n s , 8 Completeness condition, 54

Fourier -en& of, 71 Complete systems, 54-60 Antinode, 272 criterion for, 58-60 A p p r ~ h t i m offhdom b p ~ l y n ~ m , propa th of, 57-58

1aO-122 Convergence in the mean, 55 Arithmetic means: method of, 156162

Convolution of two Mans, 1% Comers, 18 Cosine series, 23

B Cosine transform, 192

Basic trigonometric system, 10,41,175-177 c ~ m p ~ of, 117-118

important comequences, of, 119-120 D

in two variables, 175-177 Derivative, left-hand, 73 Bemstein's theorem, 154 right-hand, 73 Bessel's equation, 197 Discontinuity of the &st kind, 17

g e n d solution of, 203-204 Discontinuity of the second Kid, 17 pmmetrk form of, 215

Bessel fiurctio~~~, 197-2#) asymptotic formulas for, 208-213 E evaluation of integrals involving, 218-220 of the first kind, of half-in- order, E i m i o n method:

207-208 applications of, 268-31 8 of negative order, 202-203 theory of, 245-267 of nonnegative order, 198-201 ~ u t l c t i o n s :

of the second kind, 203-205 definition of, 247 orthogonality of, 216218 orthogonality of, 251-254 relations between, 205-207 Eigenvalues : zeros of, 213-215 definition of, 247

Bessel's inequality, 54,174 existence of, -251 consequence8 of, 66,223 reality of, 25 1,253

Boundary conditions, a 5 ff. sign of, 254-255 Boundary value problems, 247 ff. Eder's constant, 204

g a e m h d solutions of, 261-264 Euler's formula, 33 inhomogeneous, 264-266 Even extension of a function, 23

Buniakowki inequality, 51 Even function, 21 333

Page 347: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

334 INDEX

Fourier-Besd caefficients, 221 order of magnitude of, 228-234

Fourier-BeseI series, 220-244 criteria for conveqpnce of, 221-223 definition of, 221 d i i t i a t i o n of, 234-237 of fundions defined on [O,fl, 241-243 of the second type, 237-241 uniform convergence of, 225-228

Fourier coefkients, 13 ff. approximate calcu1ation of, 150-1 52 complex, 34 of f b n s of two variables, 174

Fourier integral, 180-196 definition of, 182 theorem, 182

diffmnt forms of, 189-190 proof of, 188

Fourier series (trigonometric), 1-40,66-196 addition and subtraction of, 122123 arithmetic means of partial sums of,

157-1 58 integral formula for, 158

complex f ~ r m of, 32-34 conditions for convergence of, 19,7539,

178 diff'erentiation of, 129-143 double, 173-180

with different periods, 180 evaluation of, by using functions of a

complex variable, 105-1 12 improving the convergence of, 144 integration of, 125-129 list of, 147-150 multiplication of, by a number, 123 operations on, 1 lS-154 partial sums of, 73

integral formula for, 73 products of, 123-125 summation of, 155-171

uniform convergeme of, 80-89 Fourier d e s with respect toeigenfundtions,

255 Fourier transform, 1S193

inverse, 190 Functions of period 21,35-38,94

Fourier coefficients of, 35 Fourier series of, 35

Fundamental mode, 272

Gamma firnction, 199,201-202 Gibb's phenomenon, 96 GramSchmidt orthogonalization process,

65

Harmonics, 2-6 amplitude of, 3 angular frequency of, 3 initial phase of, 3 superposition of, 7

Heat capacity, 297 Heat flow in a circular cylinder, 310-316

steady-state, 31 3-31 6 whose surface exchanges heat, 312-313 whose surface is insulated, 310-312

Heat flow in a finite rod, 296-306 with ends at constant temperatwe, 299-

300 with ends exchanging heat, 301-306 with ends at specified variable tem-

peratures, 301 with ends at zero temperature, 297-299

Heat flow in an infinite rod, 306310 Htilder (Lipschitz) condition, 40 Hooke's law, 276

Improper integrals depending on a para- meter, 182-1 85

continuity of, 183 diffmntiation of, 184 uniformly convergent, 182-185

Infinite series of functions, 9 ff. convergent, 9 differentiation of, 10 integration of, 10 sum of, 9 uniformly convesgent, 9

Initial conditions, 246 ff. Integration by parts, 8

Jump discontinuity, 17 Jump of a function, 17

Page 348: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

Lqgcndrc polyrromicrls, 65 Limit:

M-hand, 17 -+hand, 17

Linearly independent f d o n s , 64,251 LiouvW, J., 258 Localition principle, 90 Lyapunov, A. M., 119

Mean square error, 51-53 minimum of, 52-53

Membrane, 282 Modulus of elasticity (Young's modulw),

276

Nodal lines, 286-288,295-296 Nodes, 272 Norm of a function, 42, 173

Odd extension of a function, 24 Odd hnction, 21 Opedonal calculus, 192 Orth08onal functions, 12,41 Orthogod systems, 4165,172174

complete, 54, 174 examples of, 14-50 Fourier coefifcients with mpcct to, 43 Fourier series with mpect to, 43 n o r m a l i , 42,173 in two variables, 173-1 74

Orthonormal system, 42 Overtones, 273

ParseVal's theorem, 119,177 wi, 1

stpndard, 8 Periodic cxtemions, 15 ff. Mdic functions, definition of, 1 properties of, 1-2

F k d s e smooth functions, 18 Poisson's kernel, 164

Regulat points, 108 Restoring fom, 3 Riemann-Lebesw lemma, 70

Scalar pduct, 61,65 Schwarz inequality, 51,63 Separation of W b h , 246 Simple harmonic motion, 3 S b series, 24 Sine trm&orrn, 192 shgdarity, 105 Smooth functions, 18 spectral fimdon, 194 Square integrable fhctions, 50 Standing wave, 272 Steklov, V. A., 2S8 Sturm, J. C. F., 258 Sum of consines, 71 of sines, 98

Summability of Fourier series, 155471 by Abel's method, 164-170 by method of arithmetic means (Cesbo's

h o d ) , 156162 System of h i o n s :

complete, 54-60,64 complete with weight r, 256 l i l y independent, 64 vector analogy for, 6 0 6 3

Thetmal conductivity, 296 Timbre, 273 Triangle inequality, 64 Trigonometric integrals, limits of, 67-71 Trigonometric polynomials, 6

approximation of functions by, 1 1 S-117 Trigonometric series with decreasing co-

efiicints, 97-1 14 convergence of, lab105

Unbounded functions, Fourier expnnsions of, 91-94

Uniform convergeam, 9

Page 349: Fourier Series - Georgi P.tolstov (Dover 1976, 349s)

v Vibrating string, 269-275

f d vibrations of, 273-275 h e vibrations of, 269-273

Vibrations: h n c t d ~ t i c , 274 fomd, 264 f=, 264

Vibrations of a cimdar 11IRmbrane, 288- 296

c a ~ , 291-296 radial, 288-291

Vibrations of a mctmguh membrane, 282-288

Vibrations of a rod (longitudinal), 277-282 forced, -282 frts, m-280

W&ht (fhaion), 217 Wdenrtrass' approximation theorenr, 120 W e i m ' M-tcst, 10 Wmnskian, 252