You are on page 1of 745
MATHEMATICAL PHYSICS EUGENE BUTKOV St. John’s University, New York Se é 3g ee *o SryoeS ADDISON-WESLEY PUBLISHING COMPANY Realing, Massachusetts - Menle Pork, California - London - Syéney - Manila oes oe OM. Sept, “eRRS Stove’ WORLD STUDENT SERIES FIRST PRINTING 1973 A complete and unabridged reprint of the original American textbook, this World Student Series edition may be sok! only in those countries to which it is con- signed by Addison-Wesley or its authorized (rade distributors. [t may not be re-exported from the country to which it has been consigned, and it may not be sold in the United States of Americaor its possessions. Copyright © 1968 by Addison-Wesley Publishing Company, inc. All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted, in any form or by any means, electronic, mechanical, photocopying, recording, or otherwise, with- out the prior written permission of the publisher. Original edition published ‘in the United, States of America. Published simultaneously in Canada. Philippines copyright 1968 Library of Congress Catalog Card Number: 68- 11391, PREFACE During the past decade we have witnessed a remarkable increase iri the number of students seeking higher education as well as the development of many new colleges and universities. The inevitable nonuniformity of conditions present in different institutions necessitates considerable variety in purpose, general approach, and the level of instruction in any given discipline. This has naturally contributed to the proliferation of texts on almost any topic, and the subject of mathematical pliysics is no exception. There is a number of texts in this field, and some of them are undoubtedly of outstanding quality. Nevertheless, many teachers often feel that none of the existing textsis properly suited, for one reason or another, for their particular courses. More important, students sometimes complain that they have difficulties studying the subject from texts of unquestionable merit. This is not as surprising as it sounds: Some texts have an encyclopedic character, with the material arranged in a different order from the way it is usually taught; others become too much involved in complex mathematical analysis, preempting the available space from practical examples; still others cover a very wide variety of topics with utmost brevity, leaving the student to struggle with a number of difficult questions of theoretical nature. True enough, a well-prepared and bright student should be able to find his way through most of such difficulties. A less-gifted student may, however, find il very difficult to grasp and absorb the multitude of new concepts strewn across an ad- vanced text. ‘Under these circumstances, it seems desirable to give more stress to the peda- gogical side of a text to make it more readable to the student and more suitable for independent study. Hopefully, the present work represents a step in this direction. It has several features designed to conform to the path an average student may conceivably follow in acquiring the knowledge of the subject. First, the inductive approach is used in each chapter throughout the book. Fol- Towing the fundamentals of modern physics, the text is almost entirely devoted to linear problems, but the unifying concepts of linear space are fully developed rather Jate in the book after the student is exposed to a number of practical mathematical techniques. Also, almost every chapter starts with an example or discussion of elementary nature, with subject matter that is probably familiar to the reader. The introduction of new concepts is made against 2 familiar background and is later extended to more sophisticated situations. A typical example is Chapter %, where the basic aspects of partial differential equations are illustrated using the “elementary functions” exclusively. Another facet of this trend is the repeated use of the harmonic oscillator and the stretched string as physical models: no v vi PRETACE attempt is made to solve as many problems for the student as possible, but rather to show how various methods can be used to the same end within a familiar phys- ical context. In the process of learning, students inevitably pose a number of questions necessary to clarify the material under scrutiny. While most of these questions naturally belong to classroom discussion, it is certainly beneficial to attempt to anticipate some of them in a text. The Remarks and many footnotes are designed to contribute to this goal. The author hopes they answer some questions in the mind of the student as well as suggest some new ones, stimulating an interest in further inquiry. A number of cross-references serves a similar purpose, inviting the reader to make multipte use of various sections of the book. The absence of numbered formulas is intentional: if the student bothers to look into the indicated section or page, he should not simply check that the quoted formula “is indeed there,” but, rather, glance through the text and recall its origin and meaning. The question of mathematical rigor is quite important in the subject treated here, although it is sometimes controversial, It is the author's opinion that a theo- retical physicist should know where ‘he stands, whether he is proving his own deduc- tions, quoting somebody else’s proof, or just offering a reasonable conjecture. Consequently, he should be trained in this direction, and the texts should be written in this spirit. On the other hand, it would be unwise to overload every student with mathematics for two main reasons: first, because of the limitations of time in the classroom and the space in a text, and second, because physicists are apt to change their mathematical postulates as soon as experimental physics lends support to such suggestions. The reader can find examples of the latter philosophy in Chapters 4 and 6 of the text. Whether the author was able to follow these principles is left to the judgment of users of this book. Each chapter is supplied with its share of problems proportional to the time presumed to be allotted to its study. The student may find some of the problems rather difficult since they require more comprehension of the material rather than sheer technique. ‘To balance this, a variety of hints and explanations are often supplied. Answers are not given because many problems contain the answer in their formulation; the remaining ones may be used to test the ability of the student for independent work. The exercises within the text can be used as problems to test the students’ manipulative skills. For many of the methods of instruction of mathematical physics presented in this book, the author is indebted to his own teachers at the University of British Columbia and McGill University. The encouragement of his colleagues and students at St. John’s University and Hunter College of the City University of New York is grcatly appreciated. Also, the author wishes to thank Mrs. Ludmilla Verenicin and Miss Anne Marie Nowom for their help in the preparation of the manuscript. Palo Alto, Calif. EB. August 1966 Chapter 2 21 22 23 24 25 2.6 27 28 29 2.10 21f 212 2AZ 2.14 215 CONTENTS Vectors, Matrices, and Coordinates Introduction . Vectors in Cartesian Coordinate Systems. Changes of Axes, Rotation Matrices 2. Repeated Rotations. Matrix Multiplication. 2 2, Skew Cartesian Systems. Matricesin General... Scalar and Vector Fields Vector Fields in Plane. .. Vector Fields in Space . Cu ear Coordinates Functions of a Complex Variable Complex Numbers Basic Algebra and Geometry of Complex Numbers De Moivre Formula and the Calculation of Roots . Complex Functions. Euler's Formula... Applications of Euler's Formula... Multivatued Functions and Riemann Surfaces Analytic Functions, Cauchy Theorem Other Integral Theorems. Cauchy Integral Formula Complex Sequences and Series : Taylor and Laurent Series - Zeros and Singularities. . “The Residue Theorem and its Applications | Conformal Mapping by Analytic Functions . Complex Sphere and Point at Tefnity woe ee Integral Representations. woe Linear Differential Equations of Second Order Genera! Introduction, The Wronskian General Solution of The Homogeneous Equation The Nonhomogencous Equation. Variation of Constants . Power Scrics Solutions . The Frobenius Method . soe Some other Methods of Solution . vi ann nH 20 26 34 44 45 48 49 Ss 54 58 62. 7 7B 83 97 102 104 123 125 126 128 130 147 vi Chapter 4 4l 42 43 4A 45 46 aT 48 Chapter 5 SA 5.2 33 54 35 56 5.7 58 59 5.10 Chapter 6 6.1 62 6.3 64 65 6.6 6.7 6.8 6.9 6.10 Chapter 7 Td CONTENTS. Fourier Series Trigonometric Series 5 6 6 se Definition of Fourier Series Examples of Fourier Series Parity Properties, and Cosine series Complex Form of Fourier Series . Pointwise Convergence of Fourier Series Convergence in the Mean. Applications of Fourier Series . The Laplace Transformation Operational Calculus The Laplace Integral Basie Properties of Laplace Transform The Inversion Problem . ‘The Rational Fraction Decomposi The Convolution Theorem . Additional Properties of Laplace Transform Periodic Functions. Rectification . ‘The Mellin Inversion Integral . Applications of Laplace Transforms . n Concepts of the Theory of Distributions Strongly Peaked Functions and The Dirac Delta Function Delta Sequences . ‘The 3-Caleulus . Representations of Delta Functions Applications of The 5-Calculus Weak Convergence . Correspondence of Functions and Distributions | Properties of Dist Sequences and Series.of Distributions {Distributions in N dimensions . Fourier Transforms Representations of a Function 7.2 Examples of Fourier Transformations 7.3 Properties of Fourier Transforms . 7.4 Fourier Integral Theorem : 715. Fourier Transforms of Distributions.» 7.6 Fourier Sine and Cosine Transforms 7.4 Applications of Fourier Transforms. The Principe ‘of Causality . 154 135, 157 161 165 167 168 172 179 180 184 187 189 194 206 210 22h 223 226 229 232 236 240 245 250 257 260 262 266 269 27h 273 276 Chapter 8 al B2 83 84 a5 8.6 87 88 a9 Chapter 9 O11 92 93 a4 95 96 OT 98 99 9.10 OAL Chapter 10 10.t 10.2 10.3 10.4 10.5 10.6 10.7 10.8 10.9 Chapter 11 wl 1.2 11.3 114 11.5 116 17 HB CONTENTS Partial Differential Equations The Stretched String. Wave Equation The Mettiod of Separation of Variables. Laplace and Poisson Equations The Diffusion Equation : Use of Fourier and Laplace ‘Transforms ‘The Method of Eigenfunction Expansions and Finite Transforms ‘Continuous Eigenvalue Spectrum . . Vibrations of a Membrane. Degeneracy . Propagation of Sound. Heimboltz Equation Special Functions Cylindrical and Spherical Coordinates The Common Boundary-Value Problems The Sturm-Liouville Problem . Self-Adjoint Operators . Legendre Polynomials Fourier-Legendre Series Bessel Functions Associated Legendre Functions and Spherical Harmonies - Sphericat Bessel Functions . Neumann Functions : Modified Bessel Functions - Fini --Dimensional Linear Spaces Oscillations of Systems with Two Degrees of Freedom Normal Coordinates and Linear Transformations Vector Spaces, Bases, Coordinates Lincar Operators, Matrices, Inverses . Changes of Basis . {nner Product. Orthogenality. Unit ry Operators . The Metric. Generalized Orthogonality Eigenvatue Problems. Diagonatization Simultaneous Diagonalization.. Infinite-Dimensional Vector Spaces Spaces of Functions : The Postulates of Quantum Mechanics ‘The Harmonic Osc so Matrix Representations of Lincar Operators Algchraic Methods of Solution : Bases with Generalized Orthogonality Streiched String with a Discrete Mass in the Middle Applications of Figenfunctions ix 287 298 295 27 299 304 308 313 319 332, 34 337 MO 42 350 355 372 3at 388 494 40s ait 49 aA 433 ay? al a3 451 463 467 47t 476 483 488 492 495 % CONTENTS Chapter 12 Green’s Functions 121 Introduction. 2. rr) 122 Green's Function for the Starm-Liouville Operator... . «508 12.3 Series Expansions for G(xlf) . 2... 7 ee ee SIA 12.4 Green’s Functions in Two Dimensions . 5 2 5 1 1). 520 12.5 Se ee SB 12.6 Green's Functions with Refiection Properties 2. . 0... - 527 12.7 Green's Functions for Boundary Conditions . 9... . . . 331 12.8 The Green's Function Method . 2 2 1 1 we eee 86 12.9 A Case of Continuous Spectrum 2. 2. ee SB Chapter 13 Variational Methods 13.1. The Brachistochrone Problem. 2 2. 2 2. 2.) e - 588 13.2 The Euler-Lagrange Equation. 2... . . . . © ss 554 13.3. Hamilton's Principle. coe eee 580 13.4 Problems invotving Sturm-Liouville Operators... . . . 562 13.5 The Rayleigh Ritz Method So 585 13.6 onal Problems with Constraints. | | 5...) (36? 137 v ional Formulation of Eigenvalue Problems . . . - . . $73 13.8 Variational Problems in Many Dimensions . ... 37 13.9 Formulation of Eigenvalue Problems by The Ratio Method . |. 581 Chapter 14. Traveling Waves, Radiation, Scattering, 14.1 Motion of Infinite Stretched String... 2 2 we ee + 589 14.2 Propagation of Initial Conditions... co ee 582 143. Semiinfnite String. Use of Symmetry Properties . |... 5 595 144 Energy and Power Flow ina Stretched String... 1... 599 14.5 Generation of Waves ina Stretched String. . . . . . . - 603 14.6 Radiation of Sound from a Pulsating Sphere... . 2. + GIT 14.7 The Retarded Potential. coe ee AD 148 Traveling Waves in Nonhomogeneous Media... . .. @24 14.9 Scattering Amplitudes and Phase Shifts. Los OB 14.10. Scattering in Three Dimensions. Partial Wave Analysis... 5. 633 Chapter 15. Perturbation Methods 15.1 ‘Introduction . So 152 The Born Approximation - . Lo ee ee AT 153. Perturbation of Eigenvalue Problems|.) 5 ss. ss 680 15.4 First-Order Rayleigh-Schrddinger Theory... . . . + - 653 15.5. The Second-Order Nondegenerate Theoty . . . - 5 1 + + 658 15.6 The Case of Degenerate Eigenvalues. 2 - . Chapter 16 Tensors 16.1 Introduction. - ee eee TL 16.2 ‘Two-Dimensional Stresses Soe ee ke ee ee OD 16.3 16.4 16.5 16.6 16.7 16.8 16.9 16.10 16.11 Cartesian Tensors 5 1. ee Algebra of Cartesian Tensors | | Kronecker and Levi-Civita Tensors. Pseudotensors CONTENTS Derivatives of Tensors. Strain Tensor and Hooke’s Law . Tensors in Skew Cartesian Frames, Covariant and Contravariant Representations . General Tensors. . . Algebra of General Tensors. Relative Tensors. The Covariant Derivative . . Calculus of General Tensots . . Index . xi 676 681 684 687 696 700 705 TUL 7S Ry CHAPTER 1} VECTORS, MATRICES, AND COORDINATES 1.1 INTRODUCTION To be able to follow this text without undue difficulties, the reader is expected to have adequate preparation in mathematics and physics. This involves a good working knowledge of advanced calculus, a basic course in differential equations, and a basic course in undergraduate algebra. Rudimentary knowledge of complex numbers, matrices, and Fourier scries is very desirable but not indispensable. As for the subjects in physics, the reader should have completed the standard undergraduate. training in mechanics, thermodynamics, electromagnetism, and atomic physics. Despite these prerequisites, a need is often recognized for reviewing some of the preparatory material at the beginning of a text. Let us follow this custom and devote some time to the subject of vector analysis which has a bearing, in more than one way, on the material developed in this text, Of course, such @ review must be brief and we must omit all the details, in particular those involving mathematical proofs. The reader is referred to standard textbooks in advanced calculus and vector analysis* for a full discussion. On the other hand, we hope to draw attention to some interesting points not always emphasized in commonly used texts. 1.2 VECTORS IN CARTESIAN COORDINATE SYSTEMS In many elementary tcatbooks a vector is defined as a quantity characterized by magnitude and direction, We shalt see in Chapter 40 that vectors are much more general than this, but it is fair to say that the concept of vectors was first intro- duced into mathematics (by physicists) to represent “quantities with direction,” e.g., displacement, velocity, force, ete. Doubtless, they are the simplest and most familiar kinds of vectors. As we well know, quantities with direction can be graphically represented by arrows and are subject to two basic operations: a) multiplication by a scalar,t 6) addition. These operations are illustrated in Fig. 1.1. * For example, A. E. Taylor, Advanced Calculus; T. M. Apostol, Mathematical Analysis; W. Kaplan, Advanced Calculus. + Until we are ready to discuss complex vectors (Chapler 10) we shall assume that scalars are reat numbers. 1 2 VECTORS, MATRICES, AND COORDINATES 1.2 ee a" ta In many cases we can plot various vectors from a ‘single point, the origin. Then each vector can be characterized by the coordinates of its “tip.” Various coordinate systems are possible but the cartesian coordinate systems are the most convenient. The reason is very simple and very deep: The cartesian coordinates | of a pointcan serve as the components of the corresponding vector at the same time. ‘Thi is itlustrated in Fig. 1.2, where orthogonal cartesian systems, in plane and in space, are selected. Note that the three-dimensional system is “right-handed” ;* in general, we shall use right-handed systems in this book. Figure 1.4 Figure 1.2 We can now associate with a vector u (in space)a set of three scalars (ite, ty, te)s such that au will correspond to (Aux, Ay, Aue) and u -+ ¥ will correspond to (ug + de, Uy + by, Ue + 04). Note that no suck relations hold, in general, if a vector is characterized by other types of coordinates, ¢.g., spherical or cylindrical. In addition, orthogonal cartesian coordinates result in very simple formulas for other commen quantities associated with vectors, such as a) length (rhagnitude) of a vector: Nol =u = (ud + uh ay'?, b) projections of a vector on coordinate axes:f tz = wcos (u, i), uy = ucos (U, j), u, = u cos (u, k), * Rotation of the x-axis by 90° to coincide with the y-axis appears counterclockwise for all observers with z > 0. + Standard notation is used: The symbols i, j, k are unit vectors in x-, ¥-, and z-directions, respectively. The symbot (u, ¥} stands for the angle between the directions given by wand ¥, 12 VECTORS IN CARTESIAN COORDINATE SYSTEMS = 3 ) projection of a vector on an arbitrary direc- = tion defined by vector s (Fig. 1.3): OP =u, = wcosy = uy 60s (8,1) + uy €05 (6, J} + 1, 608 (6,10, d) scalar product (dot product) of two vectors: (0+) = Wwc08 (HY) = de + ydy + ids, ©) vector product (cross product): . , Figure 1.3 Wa X 1] = (pe — Had) Cade — WE + (Udy — MYODK. The important distinctive feature of the cross product is that fu X ¥] + [¥ X uw], namely, it is not commutative; rather, it is anicommutative: fe xX vy} = -[v Xo} Remark. Apart from its important physical applications, the cross product of two vectors leads us to the concept of “oriented area.” The magnitude of [u X ¥}, namely no [sin (a, 1), is equal Lo the atea of the parallelogram formed by wand ¥. The direction of [u X ¥) can serve to distinguish the “positive side” of the paraflelogram from its “negative side.” Figure 1.4 shows two views of the same parallelogram illustrating, this idea, * [u xv] Positive side Figure 1.6 Closely related to this property is the concept of a right-handed triple of vectors. ,Any three vectors u, v, and w, foken in this order, are said to form a right- handed (or positive) triple if the so-called #riple product iw X J) is posilive.* This happens if w is on the same side of the plane defined by the vectors u and ¥, as illustrated in Fig. 1.5. It is not hard to verify that ({w x ¥]- ¥) represents, in this case, the volume V of the parallelepiped formed by the vectors u, ¥, and w. * These vectors form a left-handed (negative) triple if (tu X v1 w)< 0. 4 VECTORS, MATRICES, AND COORDINATES 13 Exercise. Show that v = \(a X 1) -W)] under any circumstances. Show also that the sign of the triple product is unchanged under cyclic permutation of u, ¥, and w, that is, {la X v]- 9) = (lw Xu} y) = (ly X wu). Figure 1.5 13. CHANGES OF AXES, ROTATION MATRICES We have seen that a given vector wis associated with a set of three numbers, namely its components,* with respect to some orthogonal cartesian system. However, it is clear that if the system of axes is changed, the components change as well. Let us study these changes. Consider, for vectors in a plane, a change in the system of axes which is pro- duced by a rotation by the angle 6, as illustrated in Fig. 1.6. The old system is (x, y) and the new system is (x’, y’). Since u = uzi + u,j, the x’-component of u is the sum of projections of vectors wei and u,j on the x-axis, and similarly for the p’-component.{ From the diagram we see that this yields jub = upcos@ + tysing, fy = —uzsin@ + ay cos. It is instructive to note that the angle between the x‘- and y-axes is (x/2 — 6) while the angle between the y’- and x-axes is (x/2 + 8). In view of sino = cos (5 ~ 2) / ~sin ¢ = vos (3 40): we see that all four coefficients in the above equations represent cosines of the angles between the respective axes, Let us how turn to the three-dimensional case. Figure 1.7 represents two orthogonal cartesian systems, both right-handed, centered at O. It is intuitively clear that the primed system can be obtained from the unprimed one by the mo- tion of a “rigid body about a fixed point.” In fact, it is shown in almost any text- book on mechanics} that such a motion can be reduced to a rotation about some axis (Euler’s theorem). and i * Instead of “components,” the lerm “coordinates of a vector” is often used (sex also Section 10,3). + For example, Goldstein, Classical Mechanics, Section 4.6. 13 CHANGES OF AXES. ROTATION MATRICES 5 Write a = ust + uj + uk, collect contributions to ui from the three vectors w.i, uj, and u,k, and obtain i, = uy cos (i', i) + u, cos (¥, §) -F u.cos @, k), where i’ is, of course, the unit vector in the x’-direction. Note that the cosines involved are the directional cosines of the x'-direction with respect to the unprimed system or, for that matter, the dot products of i” with f, j, and k. Figure 1.7 It is clear that similar formulas can be written for uf, and ui. At this stage, however, it is very convenient to switch to a different notation: Instead of writing (tz, yy te), let Us write (15, ue, 43) and similarly (4, 4, w4) for (x, af, 2). More- over, denote by cnn the angle between the mth primed axis and the nth unprimed axis (three such angles are marked on Fig. 1.6) and by dna the corresponding cosine (that is, @mn = COSamn). This new notation permits us to write the transformation formulas in an easily memorized pattern: wy = ayitt + ayaa + A133, 4 = Gait + agaitg + aastts, Wh = Gait + dgaits + agate, or, if desired, in the compact form thn 3 D Gantt, = 1, 2,3). ‘+ From this analysis we conclude that the new components (u/,, w), 1%) can be obtained from the old components (1;, tv, va) with the help of adne coellicients. 6 VECTORS, MATRICES, AND COORDINATES 13 These nine coefficients, arranged in the self-explanatory pattern below are said to form a matrix* We shall denote matrices by capital letters. Columns Ist 2nd 3d Ist And > Rows ard Matrix A has three rows and three columns; the individual coefficients dmx are referred to as matrix elements, or entries. It is customary to use the first subscript ~ (min our case) to label the row and the second one to label the column; thus the matrix element aj: should be located at the intersection of the kth row with the ith column. The set of elements in a given row is very often called a row vector and the set of elements in a column, a column vector. This nomenclature is justified by the fact that any three numbers can be treated as components of some vector in space. However, at this stage it is worthwhile to make a digression and establish a geo- metric interpretation for the column vectors of 4. The reader should pay par- ticular attention to the argument because of its general significance. Let us imagine a unit vector u. Suppose the unprimed system was oriented in such a way that u was along the x-axis. Then the components of w are (1, 0, 0) and u actually coincides with the vector i, If the coordinate system is now rotated, the new components of u are given by uy = aul + 420 + 4130 = air, ail + 4200 + aoa = aor, & El uh = agi! + 4320 + 4330 = aor. We sec that the first column vector of matrix A is composed of the new components of vector u., In other words, we can say that the new components of i are (G11, 421, 431) and we can write f= ayyi' + aos’ + aaik’. Similar statements relate j and k to the second and third column vectors of A. Note that in this discussion the unit vectors J, j,k assume a role independent of their respective coordinate axes. The axes are rotated but the vectors i, j,k stay in place and are then referred to the rotated system of axes. ' * More precisely, a 3. 3 matrix is formed. The reader can easily construct-an analogous 2% 2 matrix to account for two-dimensional rotations. 13 (CHANGES OF AXES. ROTATION MATRICES 7 Exercise. Establish the geometrical meaning of the row vectors of matrix A representing. a rotation. The definitions introduced above allow the computation of (ui, #2, 5) from (4, 412, us) by the following cule: To obtain uf, take the dot product of the kth row of matrix A with the vector u, as given by the triple (1, 2, us). Since in this process cach row of the matrix is “dotted” with (u1, w2, 3), We may regard it as some kind of multiplication of a vector by a matrix. In fact, this operation is commonly known as vector-matrix multiplication and is visually exhibited as shown: a1 412 13 my cH G21 pz G25 I ta = oe) 931 a2 Ang uy ty Matrix A Column vector uw — Column vector u’ As we see, the old components are arranged in a cohumn which we shall denote by u. This column vector is multiplied by the matrix A and this results in another column vector, denoted by a’. The multiplication means, of course: Form the dot product of the first row of A with u for the first component of u'; then form the dot product of the second row of A with u to get u’, and similarly for us. The entite procedure is symbolically written as Au = wf. Remark, Note that the set (1, 2, #3), arranged ina column, has xof been denoted simply by u but rather by a new symbol «.* The point is that in the context of our problem, both w and w’ represent the same vector u, but with respect to different systems of axes, We must think of uw and u’ as two different representations of u, and the matrix A shows us how to switch from one representation to another, Before we discuss further topics involving matrices, let us record the fact that our matrix A is not just a collection of nine arbitrary scalars. Its matrix ele- ments are interdependent and possess the following properties. a) The columns of A are orthogonal to each other, namely, 11012 + A21a2 + agiaa2 = 0, Fy 2019 + Ay2023 + ay2M95 = 0, 413411 + A23da4 + 439031 = 0. This property follows from the fact thal the columns of 4 are representations (in. the new system) of the vectors i, j, and k and these vectors are mutually orthogonal. * The symbol should not be confused with ful, the magnitude of vector u, which is also denoted by # (p. 2). 8 VECTORS, MATRICES, AND COORDINATES 14 b) The columns of 4 have unit magnitude, namely, ai, + hy + ahi = 2 2 2 a2 + a22 + age a}y + 033 + a9 = I, 1, 1, because i, j, and k are unit vectors. ) The rows of A are also mutually orthogonal and have unit magnitude, This is verified by establishing the geometrical meaning of the sow vectors of A.* Matrices satisfying these three properties are called orthogonal, We may then conclude that the matrices representing rotations of orthogonal cartesian systems in space are orthogonal matrices. Remark, There are orthogonal matrices which do not represent rotations. Rotation matrices have an additional property: their determinant (i.¢., the determinant of the equations on p. 5) is equal to 4-1. Otherwise, orthogonal matrices can have the deter- minant equal to —1. The point is that a rotation must yield a right-handed triple (Wf, 4) of unit vectors since (i, j,k) is a right-handed triple.t 1.4 REPEATED ROTATIONS. MATRIX MULTIPLICATION ‘The matrix notation introduced in the preceding sections is particularly useful when we are faced with repeated changes of coordinate axes. In addition to the primed and unprimed systems related by a matrix A, let there be a third double- primed system of axes (x”, y", 2!) and let it be related to the primed system by a matrix B: bit baa bis B=| ba boa bes bs: baz bas Evidently, the system (x”, y’”, 2") can be related directly to the system (x, y, z) through some matrix C and our task is to evaluate the matrix elements Cpa in terms of matrix elements dan and ban. We have ‘ tf = Brite + Brats + Oras, bast + Booty + bests, Bayh + buat + Bsatlss ug f PA and , w= ayy + yatta + aisua, Gays + Gyaite + oats, i ty = Agatty + Asatte + dgalta, ub * Htwill be shown in Chapter 10 that any NX N matrix which satisfies (a) and (bo) must also satisfy (¢). +See Problem 4 at the end of this chapter. 14 REPEATED ROTATIONS, MATRIX MULTIPLICATION 9 from which it follows that = Gra + byzan + brsasiuy + Grsare + br2a2a + bisag2)ea + Guiaig + Byzaa3 + Biaaaz)us, . 1 = (boiay1 + Drager + bastas dus + (berda2 + baeGea + basago}te + (621013 + be2deg + Bogays)ua, Wy = (bards, + baade, + basagy)ur + (barare + byzdez + bazaya)iea + (631413 + bs2az3 + B33a53)e5- The maze of matrix elements above becomes quite manageable if we observe that every coelficient associated with », is a dot product of some row of matrix B and some column of matrix A. A closer look at these relationships leads us to the following statement: The element can of matrix C is obtained by taking the dot product of the mth row of matrix B with the nth cohann of matrix A. Now, if we record our relationship in the vector-matrix symbolic notation ws Au = Bt, = CH, then we are naturally led to the relation uw" = Cu = B(Au). It seems reasonable now to define the product of two matrices, like B and A, to be equal to a third matrix, say C, so that the above relationship may also i * be written as’ a? = Cu = (BA. In this sense we write Cin Ciz Cis bit bia das Min 42 413 21 C22 Cas |= | bor dae bas |-| Ger Gon aon 31 Caz C33 531 baa baa 431 39 aan or, symbolically, CaBA given that the matrix elements of C are defined by the rule quoted above. Having introduced the notion of matrix multiplication, we are naturally interested in determining whether it has the same properties as the mulliplication of ordinary numbers (scalars). A simple check shows that the associative Jaw holds: If we multiply three matrices 4, B, and C in that order, then this can be done in two ways: ABC = (ABYC = A(BC) {where it is understood that the operation in parentheses is performed first). * The difference is, of course, that in B(Ax) the column vector w is first multiplied by A, producing another column vector which is, in turn, multiplied by B. In (BA)« the matrices are being multiplied first, resulting in a new matrix which acts on x. 10 VECTORS, MATRICES, AND COORDINATES 14 Exercise. Verify this statement. [Hint: If AB = D, then the elements of D are given by 2 dren = Dy Amid in: fi Develop the matrix elements of (AB)C and A(BC) in this fashion and verify that they are the same.] On the other hand, matrix multiplication is not commutative: AB ¥ BA, and, furthermore, there is no simple relation, in general, between AB and BA. This noncommutativity feature precludes the possibility of defining “matrix division.”* However, it is possible to talk about the inverse of a matrix and this concept arises naturally in our discussion of rotations. Indeed, if we rotate our orthogonal cartesian system of axes, the new coordinates of vector u are obtained from the vector-matrix equation w= An. Suppose now that we rotate the axes back to their original position. The new components of vector v are given by (1, ta, #g) and the old ones are given by (u4, ub, 9); these components must be related by some matrix B: u= Bu, Combining these relations we obtain u= B(Au) = (BAU. Therefore, the matrix BA must transform the components (a1, ta, ug) into them- selves, It is easy to see that this task is accomplished by the so-called unit matrix (“identity matrix”) Exercise. Show that if 2 = (BA)u is to bold for an arbitrary vector u, then BA must necessarily be of the above form. In other words, the identity matrix is unique. It is now customary to call B the inverse of matrix A and to denote it by symbol AW" 50 that we have A714 = F. Since we could have performed our rotations in reverse order, it is not hard to see that AB = J as well, that is, AT1A =) AAm* and A = B~'. While two rotation matrices may not commute, a rotation matrix always commutes with its inverse. * If we write 4/B = X the question would arise whether we imply A = BXor A = XB. + See Section 10.4 for a general statement to that effect. 15 SKEW CARTESIAN SYSTEMS. MATRICES IN GENERAL i It may now be of interest to relate the elements of matrix B to those of matrix A. To- obtain bn», we should, in principle, solve the equations on p. 5 for tty, tz, ua. However, in the case of roiations, we have a much simpler method at our disposal. Let us write the matrix equation BA = J in detail: Oy, biz Ora iy G12 G13 loo ber bea bas f-| 21 azz a2g |=] O 1 0 Bs, baz bay 431 G32 a3 9 To get the first row of J we must take the dot products of the first row vector of B with each of the column vectors of A. However, we know that the latter are just the vectors I, j, k in new representation. We sec that the first row vector of B is orthogonal to J and k and its dot product with iis unity, Consequently, it could be nothing else but the vector i (in new.representation, of course), and we conclude that by) = 414, O12 = gy, and by3 = ag). Repeat this argument for the other rows of # and deduce that the rows of B are the columns of A and vice versa. This is also expressed by the formula Bn = Gam Any two matrices A and B satisfying these conditions are called transposes of each other and are denoted by B = AT and A = BT, While it is not, in general, true that the inverse and transpose of a matrix are identical, this rule holds for rotation matrices and is very useful. 1.5 SKEW CARTESIAN SYSTEMS. MATRICES IN GENERAL If the coordinate axes in a cartesian system form angles other than 90°, we have a skew cartesian system, Figure 1.8 shows two such systems, one in plane and one in space, along with the decomposition of a vector into its respective components, Figure 1.8 x 12 VECTORS, MATRICES, AND COORDINATES 15 Skew systems are specified by the angles between the axes (one in plane, three in space) which may vary between 0° and 180°, As before, &, §, and k will be used to denote unit vectors in the direction of axes. Note that we can still talk about right-handed systems in space, where the vectors i, j, k (in that order!) form a right-handed triple. The vectors are added and multiplied by scalars according to the same rules that were stated before. However, the length of a vector is now given by a different formula. For instance, for a plane vector we have from Fig. 1.8(a), by the cosine theorem, l hat” so that fot = Wid + ud + 2uguy, cos 6. In general, the dot product is no longer given by the sum of the products of com- ponents, but by a more complicated formula. As a matter of fact we shall even introduce a new name for it and call it the inner product of two vectors which is then defined by @y= The reason is that we would like (to retain the name dot product to mean the sum of the products of components of two vectors, regardless of whether the axes are orthogonal or skew.* * "The derivation of a formula for inner product in a skew system is greatly facilitated by its distributive property, namely, @-@ +) = @- y+ GW). Indeed, from Fig. 1.9, no matter which coordinate system is used, we see that (u-@ +) However, MN - MP + PN; since MP is the projection of v on u, we have ful - (4P) = fu} - vf cos (u, ¥), and similarly for PN, establishing the result. Note that this argument is also valid for vectors in space. Now we can writet for two vectors in a plane u = ui -+ aj and v= vad + vy; (uv) = (tak Bal) (ead Od) + Gayd vat) + Crd Dad = Udy + Upby COP $ Myre COS H+ Uy. = u2 + ub — Duigiy cos (e — §) = ub + uy + 2uzH, cose + fy] - cos (a, ¥). - GAN). : [fv +] cos @ = Note that this formula reduces to the usual dot product when ¢ = 90°. | * With this distinction, the dot product becomes an algebraic concept (referring to the ' components) while inner product is a geometrical concept, independent of the coordinate system. The two are identical provided the system is orthogonal. FSince (a-¥) = (vw), the second distributive law ((u + +9) = GW) +O #) is uriviat. LS SKEW CARTESIAN SYSTEMS. MATRICES IN GENERAL 13 There is no difficulty now in establishing other formulas for skew systems, in plane or in space. We shall not go into these details but rather consider another important question: the transformation of coordinates from an orthogonal to a skew system of axes. Consider, for instance, Fig. 1.10; here a skew x’y’-system with unit vectors i and j’ is superimposed on an orthogonal xp-system with unit vectors i and j. A given vector u can be represented either as u — ui + u,j or as Figure 1.9 Figure 1.10 Although i’ = 1, the components wu, and #, are not equal; rather, we have = OF = 00 - FB up — My tan y. Also, u, = OF = ny sec. We see that the new components (uz, ui) are linearly related to the old components (t,, a) and this relationship can be represented by means of vector-matrix multiplication: cA 1 -tany te This can be written symbolically as Ww = Aa, where w’ stands for the column vector (ui, ui), and u stands for the column vector (Uz, ty). The obvious difference from the previous cases is that the matrix A is no longer orthogonal. Its inverse A~? is readily calculated by solving for t, ty in terms of i, uf, and it reads 1 siny 0 cosy Note that it is no longer the transpose of A. 14 VECTORS, MATRICES, AND COORDINATES 1.6 It is still true that the columns of A are the old unit vectors in new representa- tion (Fig. 1.11), that is, i Li +0-7, j= —tany-¥ + secr-j’ The rows of matrix A do not have a simple geometrical interpretation, but the columns of matrix A—? do have one.* From the above analysis we may conjecture that, in general, a change from one set of unit vectors to another in a plane or in space involves a linear relationship between the old and new components of a vector, expressible by a vector-matrix multiplication uw = Au (Aisa 2 X 2 or a3 X 3 matrix). We shall consider this problem in detail in Chapter 10. For the time being we shall mention the fact that not every matrix A can represent such a relationship. Consider, for instance, the following hypothetical relation between the new and old coordinates of some vector u: ue = du, — Quy, Wy = Dee — ty. If we attempt to solve these two equations for u, and , we find that they have #o solution if ul and uy, are arbitrary. We say that the matrix 4-2 2-1 Az does not possess an inverse; such matrices are called singular matrices, It is not difficult to see that Figure 1.11 in this example the pair uZ, u, cannot possibly repre- sent an arbitrary vector in plane: Our equations imply u; = 2m), so that there is only one independent component instead of the two required for a plane. Remark. It is of interest to note that if = 2uj is actually satisfied, our system of equations has an infizity of solutions for uz and ny (since two equations reduce to a singte one), Furthermore, if an additional requirement 2, — 2u, is imposed, the system has a unique solution (it — 4e3 ty — 4ul). All these features should be remembered since they are important in physical applications. 1.6 SCALAR AND VECTOR FIELDS So far we have been discussing constant vectors, but we can also contemplate vectors which depend on one or more variable parameters. Fhe simplest example ! is, perhaps, a position vector which depends on time f. In a fixed coofdinate systein, this is equivalent to saying that its components are functions of time * If A were orthogonal, the rows of A would be identical with columns of A~! (see pp. LI and 440). 1.6 . SCALAR AND VECTOR FIELDS 15 and we write w= uf) = u(Q-i+ wy) + ut) k, Such vectors can be differentiated with respect to the variable f according to the definition 4 # ate) = tim L490 — o), Ano at With u(/) and u(t + Ad) expressed in terms of their components, it is trivial to deduce that 2uy-% dig 5 +24 ding so that the operation of differentiation of a vector is reduced to differentiation of its components. While vectors depending on time are widely used in mechanics of particles, we shall be more interested in another type of variable vectors: those depending : on space coordinates (x, y, 2)-* Such vectors are said to form vector fields and can be denoted as follows: . = U(x, yy Z) = uals Ys Zi yl, Ys ZI bax, ¥, 2K. Common examples are electric and magnetic fields in space, velocity field of a fluid in motion, and others. The simplest kind of such a field is probably the so-called gradient field{ which can be derived from a single scalar function (x, », z), usually referred to as a scalar field. Familiar cases of scalar fields include the temperature distribution in a solid body, density of a nonhomogeneous medium, clectrostatic potential, etc. A scalar field gives rise to numerous other quantities through its various partial derivatives. Jn particular, let us concentrate our attention on a) the total differential ag ax Be gy 4 28 do = Sede + So dy + 3 a, and b) the directional derivativet de _ ag dx , de dy , dg dz. ds ax ds ' ay ds | az ds * These vectors may also depend on time, but we shall be mostly interested in instanta- neous relationships, where ¢ has some fixed value, t This is also called conservative field or potential field, Cee { Rate of change of ¢ per unit length in some particular direction characterized, 3yy fy fem the clement of are ds of some curve. See, e.g., Apostol, p. 104. ie 16 VECTORS, MATRICES, AND COORDINATES 1.6 ‘The expressions on the right-hand side of the equations in (a} and (b) have the appearance of a dot product. It is convenient to define the gradient of a scalar field (x, y, z) by the vector ay, -%; ae grade = Seat eit Sek Then we can write # = (grad ¢- 80), dp = (grad e- ds) and where ds = dxi+ dyj+dzk represents infinitesimal displacement in some direction and ax diy So = i+ Ex is the unit vector in the specified direction * Since every differentiable scalar field generates a gradient field, it is natural to ask whether any given vector field u = u(x, y,z) may not be the gradient of some scalar g. The answer is negative and this becomes clear as we examine the basic properties of gradient fields. In this survey we shall need certain assumptions regarding the differentiability of various functions and analytic properties of curves and surfaces involved in vector analysis. We shall mention these assump- tions as we need them. In many cases they can be relaxed and the results can be generalized, but we shall confine ourselves to the common situations encountered in physics. A curve in space is called smooth if it can be represented by x= x), ys Wh 7 = 2), where x(#), y(1), and z(t) have continuous derivatives with respect to the parameter ¢ (for a curve in a plane, simply set z = 0), Smooth curves possess tangents at all points and a (vector) line element ds can be defined at any point. The smooth- ness also guarantees the existence of line integrals} This last property is trivially extended to piecewise smooth curves; ic., those consisting of a fitite number of smooth parts. We shall assume that all curves considered by us are piecewise smooth. / Regarding the differentiability of various functions, we must remember the following definitions and statements: the interior of a sphere of arbitrary radius ¢ (usually thought to be small) centered at some point M(x, y, z) is called a neigh borhood of this point (in a plane, replace “sphere” by “circle”). Ifa set of points — : * Observe that dx/ds = cos (So, i}, etc., are the directional cosines of the ‘direction defined by ds or by S9. . + We shall assume that all integrals are Riemann integrals which are adequate for our purposes. For example, see Apostot p. 276. {A more precise term is an eneighborhood. 16 SCALAR AND VECTOR FIELDS = 17 is such that it contains some neighborhood of every one of its points, then it is called an open set. For instance, the interior of a cube is an open set; we can always draw a small sphere about each interior point which will lie entirely within the cube. However, the cube with boundary points included is no longer an open set. The reason these concepts are needed is that partial derivatives of a function in space are defined by a limiting process that is ticd to a neighborhood. We must make sure a region is an open set before we can say that /(x, y, z) is differentiable in this region. In addition, we shall be mostly interested in connected open sets, or domains. ‘These are open sets any two points of which can be connected by a polygon, ie., a curve which is formed by a finite number of connected straight-line segments. From now on we shall assume that all our piecewise smooth curves lie in domains where the functions under consideration (scalar fields and components of vector fields) possess continuous first-order partial derivatives. Let us now return to the properties of gradient fields. Suppose that u = grad (x, y, z) and consider the following integral between points Mx, YonZ0) and —-N(x4, ¥15 Zi) taken along some curve C: L (u-ds) = L@ e+ 3 dy + ean). Along Along ¢ Using the parameter # as the variable of integration, we have Nn t. de) = ag dx | dy dy | ay de [* de _ f (ds) -f. (e a+ 3y det az di dt dt = en) — olta), Along C where f) and #, are the values of the parameter 1 corresponding to points M and N. We see that the integral is simply the difference of values of g(x, y, z) at points Nand M and, therefore, is independent of the choice of curve C. Conversely, if the integral fr (u- ds) is independent of path,* then keeping Mf fixed and treating W as a variable point, ' we can define a function oles 2) = FP Gu ds) mf" Gan do + ty dy + 4 * From now on, we shall occasionally use the term “path” to indicate a piecewise smooth curve, 18 VECTORS, MATRICES, AND COORDINATES 1.6 Path C, Mt 2 N Path Cy Path C Path ue ‘ N (a) ) Figure 1.12 It is now simple to show that grad ¢ = u. For instance, + Azw.2) oe +88, 7,2) — 968952) = (OOP Gas de + ay dy + oe de) Key) 24 Bey. a fT dy Htae2) and the statement uz — dy/dz follows from the fundamental theorem of integral calculus. ‘We have then established the following theorem: The necessary and sufficient condition that u = grad ¢ is the independence of path of the integral { (u- ds). An alternative way of stating this result follows [rom consideration of the integral ° § a -as) over a simple closed path C, called the circulation of vector u around C. By simple closed path we mean a closed path which does not intersect itself * The following theorem holds: The circulation of u vanishes for an arbitrary simple closed path C (in a domain D) if and only if the integral [%, (a - ds) is inde- pendent of path (in D). Indeed, let C (Fig. 1.122) be a simple closed path. Choose two arbitrary points M and N on C and write (weds) = f (as) +f" (ads) = f" (ads) — f” (wd), ¢ Mw Nv Mf Me Along Cy Along Cs Alone Ch Along Cz If the integral fi (u - ds) is independent of path, the right-hand side vanishes and the cireulation is zero. Conversely, if two paths C, and Cz connecting two points (Fig. 1.12b) do not intersect (in space), a simple closed path can be formed from them and the above equation holds. If the feft-hand side is zero, so is the right-hand side, yielding the independence of path. If C; and Cy intersect a finite number off times, the proof is obtained by splitting the closed path into a finite number of simple * This property permits us to assign the direction of integration around the curve, char- acterized by the vector ds, in a unique fashion. Lé SCALAR AND VECTOR FIELDS 19 closed paths. {n the rather excepti an infinite number of times, a Ii case to the preceding one.* al case when Cy and Cz cross each other ing process can be invoked reducing. this Remark, Within the context of the above discussion, it is emphasized that by a(x, y, z) we mean a well-defined single-calued functiont over the entire domain and nothing short of this requirement wilt suffice. In many treatmentst of the magnetostatic field H, one introduces the so-called scalar magnetic potential X so that H = gradX and yet the circulation of H does not vanish over some contours. However, in all such cases it is impossible to define X uniquely over the entire contour (for those contours for which it is possible, the circulation of H does indeed vanish). It should now be clear that many vector fields do not fall into the category of gradient fields since it is easy to construct a vector u for which the integral J(u+ ds) will actually depend on path. It is perhaps even easier to sketch some such fields, a task facilitated by the introduction of the concept of field lines. These field lines are curves with tangents directed along the vector field u at every point. For instance, Fig. 1.13 shows the velocity field (in the plane) of a fluid rotating around a circular obstacle. In this case the field lines are the tra- jectories along which the particles of fluid actually move. Figure 1.13 Figure 1.14 It is evident that the circulation of the velocity vector u around any one of the circles in Fig. 1.13 cannot be zero (the product u - ds has the same sign at each point of the circle). Consequently, the above field cannot be a gradient field. ‘The velocity field of a fluid is, perhaps, the best starting point for investigation of other types of vector fields because it naturally jeads us to another fundamental concept: the flux of a vector field. Consider the element dS of a surface S (Fig. 1.14). Just as in the case of curves, we shalt deal only with piecewise smooth surfaces, i.e., those consisting of * The details may be found in O. D. Kellogg, Foundations of Potential Theory. { A multivalued function is not one function, but a collection of severai different functions. For example, Reitz and Milford, Foundations of Electromagnetic Theory, Section 8.8. 20 VECTORS, MATRICES, AND COORDINATES 7 a finite number of smooth portions. By a smooth surface we mean a surface representable by x= xg. y= Weg 72 = 27,9), where p and g are independent parameters and the functions x, y, and z have con- tinuous first partials with respect to p and q in the domain under consideration. Smooth surfaces possess tangential planes at all points and can be oriented; that is, we can distinguish between the positive side and the negative side of the surface. We shalt also assume that our piecewise smooth surfaces are constructed in such a way that they are oriented.* It is customary to represent surface elements dS by vectors dS of magnitudes that are directed along the positive normal to the sur- face, as illustrated in Fig. 1.14. Suppose that the vector field u represents the velocity of a moving fluid. Jt can be secn that the inner product (u- dS) repre- sents the amount of fluid passing through dS per unit time. Indeed, the particles of fluid crossing dS at time ¢ will occupy the face ABCD of the shown parallelepiped at time ¢+ dt and all particles of fluid which have crossed dS between ¢ and t + dé wilt be located at + dt inside the parallelepiped. Consequently, the amount of fluid passing through dS in the interval dt is given by the volume of the parallelepiped, equal to 5+ |ul-dicos@ = (u: dS) dt. Divide by at and obtain the desired statement. By analogy with these observations, we define, in generat, the flux of a vector field u through a surface S by the surface integral & = ffu- ds). In this formula, S can be either an open or a closed surface. A very familiar case of the latter is found in Gauss" theorem in electrostatics. 1.7 VECTOR FIELDS IN PLANE According to the material of the preceding section, integrals representing circula- tion and flux are important in the study of vector fields. For vectors in a plane, the circulation integral has the form $, (u- ds) = $ (ie dx + tty d,). Integrals of this type can be analyzed by means of Green's theorem: If C }s a (piece- wise smooth) simple closed curve in a simply connected domain D and if Poy) * For details, consult, eg., Kaplan, p. 260 et seq. . ~ + Unless stated otherwise, it is conventional to take the direction of integration in such integrals, Le., the orientation of ds, as countercfockwise. 17 VECTOR FIELDS IN PLANE. 21 Simply connected domain Doubly connected domain Figure 115. and Q(x, ») have continous first partials in D, then fear om-{f (2% | where S is the area bounded by C. ‘The importance of the requirement that Cis a simple closed curve (see p. 18) lies in the fact that we can distinguish the interior of the curve from its exterior by the following rule: As we proceed along the curve in the direction of ds we designate the region appearing on our right as the exterior and that on our left as the interior. If the curve crosses itself such a formulation leads to contradiction as should be obvious by considering a curve in the shape of a “figure eight." A domain in a plane is said to be simply connected if every simple closed curve in it has its interior inside the domain as well. Figuratively speaking, a domain is simply connected if it has no “holes” (Fig. 1.15). Without going into mathematical details we shall sketch a possible method of proving Green’s theorem which greatly facilitates its physical interpretations. First of all, note that P(x, y) and Q(x, y} can always be treated as the components ux, y) and u(x, y) of some vector field, and we shall adopt, for convenience, this identification. Let us now divide the area S into a network of meshes, as illustrated in Fig, 1.16(a). Taking the integrais f(u- ds) around each mesh in Figure 1.16 22 VECTORS, MATRICES, AND COORDINATES 17 counterclockwise direction we can easily deduce that §, Ga) = Ef way. melon (The contribution from a common boundary between two meshes cancels out because of opposite orientations of vectors ds for each mesh; this leaves only the contributions from the pieces of C) Furthermore, multiplying and dividing each term in the sum by the area AS of each mesh, we obtain fo- ds) = rfey 8) as. AW anonties Suppose now that the number of meshes is increased to infinity and that each mesh “shrinks to a point” so that AS ~> 0, If the limit lira £6 08 i as—o = {%y) exists and is independent of the shape of AS,* then the sum reduces to an integral and we have ¢, {u- ds) = fff yds. s Therefore, it remains for us to cvaluate the function f(x, y). A typical mesh is shown in Fig, 1.16(b); it need not be rectangular since the arguments presented below are valid for an arbitrary shape. If #, and uy have continuous partials, then we can write ° (udu & Gade + Me) (-ot (2), o~ con = ade + (G8) 0+ (F)0-» with the approximations being within the first order in [x — é| and |y — n)-f Here P(t, 9) is the fixed point to which AS ultimately shrinks and M(x, y) is an arbitrary point on the boundary of the mesh, Writing now f ude) = fudx t+ $y dy, : , : + Except that the largest diameter of AS must approach zero; the mesh should not become infinitesimally thia while retaining finite length. + By the theorem on existence of tolal differential, guaranteed by the continuity of partial derivatives. 7 VECTOR FIELDS IN PLANE = 23 we see that the following six integrals will be needed: fax, ¢ dy, f xdy, f va, ¢ x dx, f vay. The first two are zero, the second two are +-AS and —AS, respectively, and the last two are zero. As a result, we have me (Olly atte ¢ (ude (4H — Se) as. flo, ¥) = Kien LEG) _ Sy _ Bee, a il I Consequently, AS Ox ay where the stipulation that the partials are to be calculated at P(é, 9) can be omitted since P is now an arbitrary point within C. We have then the result Z (te de + uy dy) = f ff (tt 2) as, which is simply Green’s theorem in our notation. With regard to a vector field u = uci + uyj, the function f(x, y) is known as the cur? of u so that, by definition, £0) #. curl b= lim We have then evaluated the expression for curl u in (orthogonal) cartesian co- ordinates in plane: Ouy = du, curl uw = ‘ox oy Remark, Attention is drawn to the fact that, fora vector field in a plane, curl u is essen- tially a scalar* and not a vector. The point is that by vectors we must mean quantities expressible as ai + bj and curl is definitely not this type, whether or not we introduce the third axis. A vector field u which has zero curl at some point is called an irrotational field (at that point). If u is irrotational in a simply connected domain, then by Green's theorem, it is a conservative field (geadient field) in this domain, ie., it has zero circulation. The converse has to be worded rather carefully: If u is a gradient field (namely, u = grad y), then it is irrotational provided ¢ has continuous second- order partial derivatives. *In a more elaborate nomenclature, curl u is called pseudoscalar due to ils peculiar property of changing sign if the x- and y-axes are interchanged. See Section 16.5. 24 = VECTORS, MATRICES, AND COORDINATES 17 Figure 1.17 Figare 1.18 Example. The magnetic induction field (B-field) due to an infinite current- carrying wire is known to be (outside the wire)* B= B25, (KSA unit). a? In the xy-plane (as shown jn Fig. 1.17}, = —Bsine= —H Ym _y, B= —Bsin@ = — FE Oe B, = Boos 9 = Hol % dn EP This field happens to be irrotational everywhere except at the origin. Therefore Fc (B- ds) = Oif C does not encircle the origin, but not otherwise. A function x(x, y) may be found such that B = grad x in a simply connected domain D, but this can be done only if the domain does not contain the origin. The B-field inside the wire is known to be = tol : B30 where Ris the radius of the wire. Here Be = —(#ol/2+R*)y and By = (stol/2xR7)x. ‘The field is not irrotational and cannot be represented as grad p anywhere. FS, Let us now turn our attention to the concept of flux for vectors in plane. The obvious analog to the three-dimensional case is the integral . a= f (ung) ds =f (u-dn) . : ’ taken over a curve C (not necessarily closed) with ny being the wif normal to the curve and dn = nods. This is illustrated in Fig. 1.18 > * go is the unit vector that is tangential to the circle drawn around the axis of the wire, Le. VECTOR FIELDS IN PLANE (25 Exercise. For a flow of fluid ina plane, relate the integral & above to the amount of fluid crossing the curve C per unit time. Specify physical units of ali the quantities used. In many applications, the flux through a closed curve is involved. Evidently if ds = dxi + dy}, then* dn = dyi — xj and f, (w+ dn) = $, (ts dy ~ tty dx). Setting P = — 1, Q = tty, this integral can be transformed by means of Green’s theorem so that fo m= [Ger ae) acy provided, of course, that the partials are continuous everywhere inside C. This relationship is usually referred to as the divergence theorem (in a plane) and is written as, f,, (ada) aes au, diva = ive 4 a where is another function derivable from a vector field and is known as the divergence of u. While the above derivation is straightforward, it docs not reveal the geometric (or physical) meaning of divu. It is instructive to invoke the technique used in Green's theorem: Dividing the area S into a network of meshes, we find that it is not hard to establish that oan - DLL Mas, Alt meshes because the vectors du at the common boundary of two adjacent meshes are op- positely directed. This observation permits us to define the divergence of u as flux out of an infinitesimal area (per unit area), namely, cen = him Seca) , divu = fim, AS Exercise. Derive the formula div u = du2/dx + du,/dy starting from the above defini- tion and using the arguments analogous to those for curlu. Spell out the conditions required in the derivation. * It is a standard convention that for closed curves in a plane (and closed surfaces in space) the positive normal is the outward-pointing normal. 26 VECTORS, MATRICES, AND COORDINATES 18 Remark. From our definitions of curl # and div u itis seen that both represent a new kind of derivative, namely, a derivative with respect to infinitesimal area rather than infinitesi- mal displacement: curly = in SE, diva = i, Feu), It may be of interest to mention that the gradient of a scalar field y can be represented in a similar fashion; i.c., the following statement holds: ‘eo dn as grad yp = lim £ 450 Another interesting observation is that cur! « can be related to an integral involving da and div a to an integrat involving ds. Indeed, the identities (u> dn) = [u X ds], (a: ds) = [dn X ul, are not hard to verily provided we treat the cross product of two veciors in plane as a scalar, which is the logical thing to do.* Consequently, the following statements also hold: Fla x 4) = fldn x cutu = fim RZ, AS divu = lim aged asa Vector fields which have zero divergence are called solenoidal fields. They are very common in physics. For instance, the electrostatic field is solenoidal in the absence of charged matter; the magnetic induction field is solencidal everywhere. 1.8 VECTOR FIELDS IN SPACE We would now like to extend the analysis of the last section to vectors in space. We shall start with the flux of a vector field a through a closed surface S, becaiase this integral is, perhaps, easiest to handle. It reads a= ffoas) = ff weds. + uy dS, + te dS), where dS,, dS,, and dS, are projections of vector dS on the coordinate axes. Integrdls of this type can be handled by Gauss’ theorem: If S is a (piecewise smooth) closed orientable surface contained, along with its interior, in @ domain D and if L(x, y, z), M(x, y, 2), and N(x, y, 2) have continuous first partials in D, then OL , aM , aN $f as + Mads, + NaS) “fff (+4424) ”, ! where V is the volume bounded by S. * A typical cross product of two vectors in physics is the torque I" = [r X F] which, for vectors in a plane, is completely described by magnitude and sign (clockwise or counter- clockwise). See the remark on p. 23 which may lead to a conjecture that the cross product in a plane is a pseudoscalar (and, indeed, it is) 18 VECTOR FIELDS IN SPACE 27 Figure 1,19 Gauss’ theorem can be restated in vector notation, by identifying ZL, M, and N with the components of a vector field u = wei + 1,j + uk. The concept of divergence of u, au, diva = Fe 4 Sey ou, is teadily introduced so that Gauss’ theorem can be rewritten as ff 48) = fffeiva-av and is often referred to as the divergence theorem. The deduction of this theorem can be done again by a method analogous to that in the preceding section. In this case we cut the volume V into small pieces, say, rectangular blocks, one of which is shown in Fig. 1.19. If we calculate the flux of u through each block and add alll the results, we must obtain the flux through the outer boundary S; just as before, the flux of u through an interface between two blocks must appear in the sum twice, but with the opposite sign because of the changed direction of the “outward normal” (see the footnote on p. 25). Consequently, fp dS) = 5 fe® 9S) ay am blocks introducing, for an obvious purpose, the volume AV of each block. We now define the divergence of u by means of divu = tim Hen). avao Al Gauss’ theorem can now be deduced by ironing out the mathematical de- tails and developing the formula for divu in cartesian coordinates. We shall 28 VECTORS, MATRICES, AND COORDINATES 18 give a simplified version of this from the consideration of a rectangular block AV (Fig. 1.19). It is not hard to show that the flux of u through a rectangle such as ABCD, is given* within the first order in Ay and 4z by Pancp = (us) Ay Az, where (1t,)p, is calculated at the center P of the rectangle. Indeed, for any point within ABCD we have ‘Ou, dus ue ue + (88 pO D+ a7) pF and the last two terms yield zero when integrated over the area ABCD. Consequently, the flux of u through the face A‘B'C'D? is to be determined by the value of uz at the center of this face and is given by [woe + (te | Ay az. x Jp Similarly, the flux through the opposite face is - [ure - ae) 4 Ay Az. There is a minus sign in front because we now want the positive normal in the negative x-direction. Adding these two fluxes we obtain ou 3x) p Ox ae oF The fluxes through the other four faces are obtained in a similar way leading to the statement about the flux through a small paralielepiped AV: cw (Me 4 Oty ae tov = (4 + ay + “ur Av. ‘The expression for divergence now readily follows and concludes our analysis. Let us now consider the question of circulation of a vector in space. Such an integral has the form ¢ (u-ds) = (ide + ty dy + us de), a and we shall assume that C is a (piecewise smooth) simple closed curve in space. It nced not be a plane curve but we shafl assume that it can be spanned by a (piece- wise smooth) orientable surface 5; that is, C can serve as the boundary for such a * Assuming a positive normal in the positive x-direction. 18 VECTOR FIELDS IN SPACE = -29 Figure 1.20 surface, This situation js iWustrated in Fig. 1.20{a). It is a standard convention to define the positive direction of circulation around a surface element dS so that it forms the “right-hand screw system” with the positive normal to the surface* The positive direction of circulation around the boundary curve C is now chosen so that it coincides with such direction for the adjacent surface clement (Fig. 1.20a). When all this is done, we can quote Stokes’ theorem; If S is a (piece- wise smooth) orientable surface spanning a (piecewise smooth) simple closed curve C, mutually oriented as described, then $e», 2) di + M(x, y, 2) dy + M(x, y,2) 2] -[[(@-3 2) yaa + (22) crac +(e ae axa 5 provided L, M, and N have continuous partials in a domain containing S and C. As before, we can treat L, M, and N as the components of a vector field u. By the now familiar technique, we divide S into a network of (curved) meshes (Fig. 1.20b) and claim where AS is the area of a given mesh. As before, we calculate Eads ds) Fu, dx + ty dy + uw, dz), fim FAM eT ty OY 4830 ay AS Considering a small mesh about a point P(E, 7, 7) we have, for each point Af(x, y, 2) on its boundary T (Fig. I.2ta), Ou, (dar & de + a) @- pF ( oe) po~ a+ — ete. * A rule well known from the study of magnetic fields. 30° VECTORS, MATRICES, AND COORDINATES 1s (@) Case I () Case 11 Figure 1.21 ‘We shall need integrals like ¢, dx, 4, x dx, $, ydx, ¢, zdx, and similar ones. ‘These integrals can be evaluated by projecting our mesh on the coordinate planes. Suppose thal we want to obtain the integrals of the type Sr f(x, y)dx and $rI(% y) dy. Since the points M and M’ (see Fig. 1.21) have the same x and y, these integrals reduce to $r- f(x, pdx and Sr f(x, ») dy except for the sign* which depends on the orientation of the mesh: §. Se yax = tf. SG pdx, ete. ‘The plus sign is for Case I (Fig. !.21a), where the point M’ describes I’ in counter- clockwise direction, the minus sign is for Case 1 (Fig. 1.21b), with M’ going clockwise.f In particular, note that : $d = ¢ xd = 0 pyar = gb yar FAS’, while where AS’ is the area bounded by I”, ie., the area of the projection of the mesh * The symbol fy. by itself is meant to indicate counterclockwise integration! {see the footnote on p. 20). + The motion of M’ cannot be chosen at will because it is determined by the motion of 4 in the positive direction of circulation around the mesh. 18 VECTOR FIELDS IN SPACE 3. on the xp-plane. Since the mesh is small, we have (Fig. 1.21) AS’ & AS'cos (9, k) in Case I, AS’ = —AS cos (ng, k}_ in Case Il, (where no is the unit normal to AS). In either case, vax = —AScos (to, k). Similarly, we can deduce ¢, xdy = +AS'cos (no, k) (this integral is also needed). Other integrals are evaluated in an analogous fashion. In particular, $; 2 dx and g', x dz require projection on the xz-plane, and so on. The net result of this calculation reads Ou, ai ny a Ps ficw- ayacas| (ts — | c08 (noi) + (See Ot 205 (3) ‘au, du, | + ( -&, 008 (Mo, k) Introducing the curl of a vector in space (which is now a vector, in distinction to the plane case} by means of the relation . -ds) lim £0248) asso AS = (curl ano), we have the result a ( ‘tt. duy\, . fatty aus curl = Sie el (Be i oe) * which also gives rise to the statement that circulation of u around an infinitesimal oriented area described by dS is equal to (curlu-dS). Thus Stokes’ theorem can now be-written in the compact form:* te (w- ds) = freon u- dS). As in the case of vectors in a plane, the vector fields satisfying diva = 0 are called solenoidal and those satisfying curl u = 0 are called irrotational. The concept of irrotational field is closely related to the concept of conservative field | {a = grad y) but these two fields should not be identified because of topological complications, In particular, if a field is irrotational in a domain D, if docs not’ * The shape of the area element is irrelevant; because of this we can identify dS, in dS = dS,\ + dSyj + aS.k with dy dz, ete. 32 VECTORS, MATRICES, AND COORDINATES 18 { folfow that its circulation about an arbitrary simple closed curve in Dis zero. The point is that we may not be able to invoke Stokes’ theorem because we may not be able to construct a suitable spanning surface S which would lie inside D. Example. Consider a tightly wound coif of current-carrying wire in the shape of a torus, The B-field inside the torus is irrotational; indeed, cur} B satisfies the equation* curl = pod + conto OE (MKSA units) and there is no current density J and no displace- ment current in the interior of the torus; we assume a dc-situation. However, the B-field at the center of each turn of the coil is known to be B = nol yielding a circulation along the central cirele C of the torus: §, ds) = wont - 2nR x 0. It is easy to see that any surface spanning C must necessarily extend outside the torus and cut through the windings where curl B # 0. Since the quantities divu and curlu in a plane have been defined as area derivatives (p. 26) and since div u in space has been defined as a volume derivative (p. 27), it may be conjectured that curl u in space is also reducible to a volume derivative. This is in fact true and the formula reads = im FES X48 curl t = fim PT avo where the surface integral is over the boundary enclosing the volume AV. We shall sketch the proof of this relation by considering AV in the form of a rectangular parallelepiped as shown in Fig. 1.22. The contributions to #1dS X ul from the top face involve only u,i and #,j and yield, in the usual notation, fees x me axay (oe 2 ¥) 5 w+) i. Top face where P is the center of the parallelepiped. The bottom face will have dS with the direction reversed and [fas x oy -axay (x — 2 Im n+ oe 4. Bottom inca * One of Maxwell's equations; see, e.g., Reilz and Milford, pp. 296-297. 18 VECTOR FIELDS IN SPACE = 333. Adding these, we obtain au) 5 (Nd aya az) i — (az), | y Az. Contributions from the other four faces are treated similarly, establishing the result, Remark, The gradient of a scalar field g can also be represented as. a volume derivative, namely dg = fi Sheds | grado = tim ay We shall conclude this scction by mentioning some quantities obtained from the repeated operations involving gradient, divergence, and curl. First of all observe the identity curl grad y = 0 representing the statement that a conservative field is always frrotational (pro- vided, of course, that the second-order partials of g are continuous, as mentioned on p. 23). The second operation of similar type yields the definition of the Laplace differential operator V7 ot simply the Laplacian,* Vp = div grad g, with the well-known expression in cartesian coordinates 2. Ca Vex ax? + ay? + i Both of these operations have their counterparts for plane vectors. There are also two operations which are only possible for vectors in space: div curl u and curlcurlu. Straightforward calculation in cartesian coordinates yields div curlu= 0. A more sophisticated argument of some interest is to take a closed surface Sin a domain D where curl « is defined (and its components have continuous partials). Split Sinto two parts, 5 and Sa, by a curve C, as illustrated in Fig. 1.23. By Stokes’ theorem, f, (u- ds) = ff(curta-asy) = ~ ff teort wd); 3 Ss * The symbol V? is related to the so-called “nabla” or “del” operator a a a "ae tay + Bae capable of representing gradient, divergence, and curl by means of the notations grad ¢ = Ve, diva =s (Vu), and curl a= [VX ul], which are sometimes quite handy. ed VECTORS, MATRICES, AND COORDINATES 19 the minus sign arises from the “wrong” oricntation of S; with respect to C. Consequently, (curl u-dS,) + f[(eurl u- dS2) = ff (curl a- dS) = 0 [fonts 4 foe crf for any closed surface in D. Applying the divergence theorem, we obtain, f[faivcunu-dy =o. v Then, since this is true for an arbitrary volume Vin D, it follows that div curl u = 0. Regarding the operation curl curl u, we can decive the following expression for the cartesian coordinates: curleurlu = grad divu — V’u, where the symbol Vu stands for the operation V?ud + V7uyj + Vuk. This formula, however, is not valid for other coordinate systems.* Exercise. Verily the above formula using the expressions for divergence, gradient, and curl in cartesian coordinates. 19 CURVILINEAR COORDINATES Sometimes it is more convenient to use coordinate systems other than cartesian. In general, a point in space can be described by three parameters which we will denote by [, 7,2. A well-known example is given by Spherical coordinates 7,6,¢, as shown in Fig, 1.24, along with the usual cartesian coordingtes x, y, z * Of course, it is possible to define the operation ¥?u by (grad divu — curl curly) in any coordinate system, but this is not done since V?u does not reduce to the operation “div grad” applied to components of u. 19 CURVILINEAR COORDINATES 35 which are related to 7, 6, 6 by x=rsinécos¢, y=rsinésin¢, = rosé. In general, x, y,z can be thought as being functions of 1, m, 2: x= xm, y = ylm,n), = 2hm,n). We shall assume that at least within some domain D in space, these functions have continuous derivatives and can also be solved for /, m, a: meurve t m= m0, y,2) Hx 2). MO) n as Fs 2). Observe that this implies that the Jacobian ay, y, 2) aG, m,n) oO Figure 1.25 docs not vanish." Let us now choose a particular point Mf with cartesian coordi- ’ tales (£, 4, ¢), it can also be denoted as A4(A, p, v), in terms of the coordinates iin,n. He we keep m = = const and m = y = const and change /, then we obtain a (smooth) curve passing through M which may be called the Feurve. Similarly, we can define m-curve and n-curve. T) shown in Fig. 1.25. Further- more, we can introduce unit vectors Ip, mo, and mg along the tangents to these curves (pointed in the direction of increasing /,m,1). This establishes a local system of axes. For convenience, the labels /, m,n are chosen so that (Io, to, No) form a right-handed triple. Our Iocal systems possess, in general, the following features which distinguish them from the system formed by cartesian unit vectors {, j, and k. {. The axes may not be orthogonal; moreover, the angles between the axes may change from one point to another. 2. The orientation of Io, mto, No (with respect to i,j,k) may change from one point to another, even if the angles between the axes remain the same. 3, The physical meaning of parameters J, 1,2 may not be the length, and df, dm, dn need not be identical with the elements ds of arc in the respective directions. Let us investigate properties (1), (2), and (3). We can always think of point M as being defined by a position vector + = xi-+ yj + 2k. Treating x, y, z as * See, ¢.2., Kaplan, pp. 31 ef seq. 36 © VECTORS, MATRICES, AND COORDINATES 19 functions of f, m,n we can write r= x(im, ni + (maj + 20m, rk. Changing r by dr amounts to changing x, y, z by dx, dy, dz which is, in turn, caused by changing f, m, a by df, din, da. We have the following general relations: a 4 % am 4 dx = tt ag et a = 4 Pam 4 2 dy= at ag tt + 5g = 4 Zam 4% dz= Het 5m! + an If we move along the /-curve, then da = da = 0, and dr becomes* (elon = (Be 14 dy-5+ de Wan = (Sit Bit F a. This defines the derivative of r with respect to the parameter J: ox, , ay, , a = ait ait ae By its very meaning, or/af is a vector along the direction of ly. Therefore, Io can be expressed as tp = oral _ (ax/ani_ + (ey/a)i + (z/ak__ lor/all xa? + Gy/al? + (@z/al)? The quantity ~ (2) @ f (zy = & + Ma) * Na has a simple geometric interpretation: The length of elementary arc ds produced when only / changes is given by ds = |(@t)m.nf = rdf. In a similar fashion we deduce ng = Granite (ov/am + @z/amk, ( ay ( a ry ( ay , ny = Gan + Capsan + @2/amk, 4, (#) + (zy {xy * We use the nolation familiar from thermodynamics: (dr)m,. means such dr where mr and mare kept constant. 19 CURVILINEAR COORDINATES 37, Suppose now that the triple In, mto, no is an orthogonal triple. Then we must have the relations ax ax, ay ay , az az a am * at amt at am ~ % ot These relations are satisfied for most coordinate systems employed in physics. In particular, this is true for spherical and cylindrical coordinate systems, as can be easily verified. This analysis clarifies feature (1) of local systems of axes. Regarding the orientation of local axes, note that it docs indeed vary from point to point for spherical and cytindrical coordinate systems. In fact, this is the characteristic Properly of curvilinear coordinate systems, as opposed to cartesian ones. Fealure (3) is also iNtustrated by spherical and cylindrical coordinates where some of the parameters J m,n represent angles rather than lengths, As a general rule, the elementary displacement dr decomposed along the local system of axes will read or a= Sd + © don + SE dn = hadl- Sy + bm din mtg + fg t+ Mo. Let us assume that the local system is orthogonal;* then the element of arc is given by a simple formula, ds = {del = Vibd® + am + de. For instance, for spherical coordinates, A, = 1, Ay > r, hg = resin é, and ds = Va de® Ere sin® 6de?. We shall conclude our survey of curvilinear coordinates by the derivation of formulas for common differential operations in vector calculus. In order to express grad ¢ in terms of new axes and new variables, we could start with grad p = then use de _ ap a, ap am ax = af ax t am ax t and also express |, j,k in terms of Ig, mo, Mo. A quicker way is to ulilize the statement =de= a4 oe (grado dt) = do = SF dl + SE dim + 2 dn * The associated coordinates are then called orthogonal coordinates, 38 VECTORS, MATRICES, AND COORDINATES 19 {b) Figure 1.26 and rewsite it in the form (grad pdr) = (i 9) heal + be SE) hed + (t 36h, dn from which it follows immediately* that 1 a L a L a grad ¢ = 5G Selo + 5 5m Mot 5, 5a The calculation of divergence can be also carried out starting from the general definition diva = tim HU), avso AV Without loss of generality, AV can be taken as a volume element with the sides along the /, m-, and n-curves (Fig. 1.26a). In general, the flux through an ele- mentary area oriented in the Eo-direction is given by t+ hin din - fg dit. As we subtract the fluxes through the areas M'N’P'Q’ and MNPQ we must hot forget that not only 1, but also Ay, and Ay, are functions of f,m, and xf By an argument similar to that on pp. 27-28 (we give here a simplified version), we deduce that the net outward flux through these two faces is 3 (ahha) al dem da. Adding the analogous contributions from the other four faces and dividing by the volume of our volume clement (which is fd hy dra hy dn), we oblain immediately . diva = EL? (a shatn) + 2 (Unlinks) + 2 (ugha J}. Hy lat ™ one” ane : * Recall that de is arbitrary; therefore, by setting dm = dn = 0 we obtain (grad ei: = (/A)Ge/2N, ete. We tacitly assume that the coordinates are orthogonal. } In-cartesian coordinates fi = fn — fn = 1 19 CURVILINEAR COORDINATES 39 Example, In spherical coordinates we identify J, m,n with r, 8, ¢, in that order. Then hy = hy = 1, lim = lig = 7, hy = hy = sin @, and 1 fa pa. a. a div w= asia g ae 17 sin Bu) + sp (rsin @ue) + 3 tus} This may be simplified and, if desirable, ultimately reduced to 1 au rt divu= a yt 244 cot 6 Lang | tut ine op r The curl of w can be deduced from the circulation of u around the facts of the very same volume element. For instance, the face MNPQ yields (Fig. 1.26b) P Ae (8A) +f (de) = Cn der — (nha ddnra = 32 Cun) dd, Pwd f way which is combined to form —(tadin EPG + Clmbim dn)aiw 1 2 ~ £m) din dt a @ (curl w)t fin dit hy dn = (2, Gruln) — 55 Cental} dm dn. This determines the Lcomponent of curl w, The complete formula reads a a cula = 7 (2 (nln) — 2 unbal lo + 5K | Rem — Scuba} mo +7 [Btu 1 2 ho} no. Finally, the expression for the Laplacian V2 is obtained by combining the formulas for gradient and divergence: Rnlin Oy fir ap 8 (htky, ag }. = div grad u = ila {5 ("ae th i) 4 3, (tt ie) + 2( Tin 2) For instance, in the spherical system this reads (after trivial simplification) ve = BEAM) + tng & (sin 0% +e, P ar F sin 6 36 FF sin? @ ag? 40 = VECTORS, MATRICES, AND COORDINATES BIBLIOGRAPHY Apostot, T. M., Mathematical Analysis, Reading, Mass.: Addison-Wesley Publishing Co., 1957. Gowpsrein, H., Classical Mechanics, Reading, Mass.: Addison-Wesley Publishing Co., 1959, Kaptan, W., Advanced Calculus, Reading, Mass.: Addison-Wesley Publishing Co., 1957. Kettoce, O. D., Foundation of Potential Theory, Berlin: Springer Verlag Obg., 1929. Tavtor, A. E., Advanced Calculus, Boston: Ginn & Company, 1955. Retrz, J. R. and E. J. Mitrorp, Foundations of Electromagnetic Theory, Reading, Mass. : Addison-Wesley Publishing Co., 1960. PROBLEMS 1. Let two vectors in a plane, uw; and uz, be defined by the polar coordinates of their tips: @1, 71) and @2,r2). If ug = ur + uz is defined by (8a, ¢3), show how @3 and 3 are related to 0), 02, 1, and rz. . A triple vector product of three vectors is defined by the expression [u X [¥ X w]}- Show that for any three vectors the following identity holds: [u x Wy x wh) + Uy X Ow ul] + fH x fo X I = 0. » Hint: Use the vector identity {a x [b x e]] = bla -c) — ea -b). ‘The above formula, known as the Jacobi identity, appears in a variety of contexts in physics and mathematics. 3. Consider the following three vectors in space given by their coordinates uf bP, VEEP w-h-BD a) Verify that these vectors are unit vectors, orthogonal to each other, and form a right-handed triple, if ordered as above. b) Construct the rotation matrix transforming the old components of a vector (namiely those with respect to 4, j,k) 10 the new ones (with respect to u, ¥, #7). ©) Evaluate, by vector-matrix multiplication, the new coordinates of the vectors 2(0, 3, 2), b(—L, 4, —3), and (2, —2, —2). Can you give a geometrical interpre- tation of the peculiar behavior of vector ¢? 4, 8) Show that the triple product (p. 3) of vectors ule, w2, ay), v(t1, 42, 23), and we(w1, 2, ws) can be expressed by the determinant = (ux vw). ~ a ™ PROBLEMS 4. b) Using this, prove that if a 3 x 3 matrix is orthogonal, then its determinant can have only two values, either +1 or — 1. ©) Consider the matrices 1 0 0 va $ 0 A=| 0 1 oO], B= $0 -v3;2 o |, o 0-1 0 0 “1 -1 0 6 2wiI/s —3Vi0 4 c=/ 0-1 of], p= : -p va |, oo 1 g 4 0 and indicate which ones represent rotations. meaning of others, Compare, in general, the ith matrix element of AA with that of BA, for 3 X 3 matrices A and B. Construct two noncommuting 3 X 3 matrices of your choice, i.e., such that AB ¥ BA. According to the discussion on p. 4, the matrix Also, describe the geometrical cos@ sin® —siné cosé@ represents a rotation of axes in planc. Show that sin 26 cos 28 cos 34 sin 38 A? = AA= and A® = AAA = —sin 28 cos 26 —sin 30 cos 3 and give the geometrical interpretation of this result, Show that the matrix does not represent a rotation of axes, Give a geometrical interpretation of matrix B. [Hint: Draw the old and the new coordinate axes, as well as the straight line y = xtan (§/2).] A2 ‘VECTORS, MATRICES, AND COORDINATES 8. Find the inverses of the following matrices by solving the equations B.A, = 7 or otherwise: 2-10 4 3 a AL= aouidl, #& 0 -UWVE v2 -3 04 wis -Vi6 ~V2/6 -3 -§4F 1 3 2 As=| -} @ 4 |, 4e= 1 ~1 o |, a 48 2 o Comment on the cases Az, Aa, and Ay. y 9. Let (x', »”) be the coordinates of a point in a skew y cartesian system in plane. Let the x'- and y’-axes make angles a and 6, respectively, with the x-axis (Fig. 1.27). Show that the equation of a circle with radius R and the center at the origin reads x? 4 y® + 2x’y' cos (8 — a) = RB? x 10, Show that the vector v = 21+ § — 6k cannot be G expressed as a linear combination of the vectors * mp =it+j+ 2k, w= 3-5, ws = Atk Figure 1.27 Show that the vector w = —2j— 3k can be expressed in this fashion, in more than one way. Give the algebraic explanation of these facts. Also give a geometrical interpretation. 1f, Evaluate the following integrals around the circle x? + y? = 1; use Green’s theorem if it is convenient: a) S(u- ds), where u = (Qy? — 3x")i + xy — 29), by) $ (2x? — y3) dx + (x? + yay. c) $(r-dn), where ¥ = (x? + y?)i — 2xyj (dn is defined on p. 24). 12, Let F(x, y) — x? — y?, Evaluate a) §8B (grad F + ds) atong the curve y = x?, b $ Fas around the circle x? + y? = 1. Here @F/dn is the directional deriva- 2 five of F along the outer normal and ds = |ds|. 13. Show that the vector field u = yzi + zxj + xyk is both irrotational and solenoidal. Find g such that grad y = u. Can you find a vector field A such that curl A = u? 14, Prove the following identities for scalar fields f, ¢ and vector fields u, v in space: a) grad (fe) = ferad y + e grad f, , ° yb) curl (fa) = fcurlu + [ered f X ul, ©} div fu X y] = (v- curla) — (a curly). 15. PROBLEMS = 43 Using divergence and Stokes’ theorems, if it is convenient, calculate the following integrals, a) $s (a- dS), where u the origin, : b) sv dS), where v = x‘ + y5} + 25k and S is the sphere as in (@), O Hs (edydz + ydz dx + z dx dy), where S is the sphere as in (a), d) $v (u- ds), where w= —3yh + 3xj + k and Dis the circte x? 4 y? — 1 lying in the plane z = 2. + ¥3j + 2% and S$ is the sphere of radius R about . A flat disk rotates about the axis norma! to its plane and passing through its center. Show that the velocity vector y of any point on the disk satisfies the equation curly = 2a, where w is the angular velocity vector. . Consider a conducting medium with variable charge density p(x, y, z) and variable current density J(x, y, 2). Let V be an arbitrary volume within the medium bounded by a (piecewise smooth) closed surface 5. Considering the total amount of charge inside V and the amount entering it per unit time through the surface S, deduce that 4 f[focsnar = - Ff o0sn2-49) ; 5 With the help of the divergence theorem, deduce the so-called equation of continuity diva + Bao, |. Using the techniques employed in Sections 1.7 and 1.8, outline the possible proofs of the following statements: lim avo AV rad y = lim, Fed" a plane), grad p = (in space). ase AS . Evaluate the quantities f,, Ay, As (See p. 36) for the cylindrical coordinate system. Using appropriate formutas from Section 1.9, write the expressions for gradient, divergence, curl, and the Laplacian ia cylindrical coordinates. CHAPTER 2 FUNCTIONS OF A COMPLEX VARIABLE 2.4 COMPLEX NUMBERS In the course of study of roots of algebraic equations and in particular the cubic equation, it has been found convenient to introduce the concept of a number whose square is equal to —1, By a well-established tradition, this number is denoted by #, and we write i? = —Iand i = /—T. If we allow ito be multiplied by real numbers, we obtain the so-called émaginary numbers* of the form bi (where bis reat), If the usual rules of multiplication are extended to imaginary numbers, then we must conclude that the products of imaginary numbers are real numbers; moreover, their squares are negative real numbers. For instance, GBA 4) = 3-4? = (-12(-D = 2, (~5i)? = (—5)*? = —25. If imaginary numbers are adjoined to real numbers, we have a system within which we can perform multiplication and division (except by zero, of course). We say that such a system is closed under multiplication and division, However, ‘our system is not closed under addition and subtraction.f To eliminate this de- ficiency, so-called complex numbers are introduced. These are numbers which are most often written in the form a+bi (a,b = real numbers) and are assumed to obcy appropriate algebraic rules. As will be shown below, the system of complex numbers is closed under addition, subtraction, multi plica- tion, and division plus the “extraction of roots” operation. In short, it has all the desirable algebraic characteristics and represents an extension of the real number system. The study of complex numbers is invaluable for every physicist because the description of physical laws is much more complicated without them. * Imaginary numbers are also called pure imaginary numbers to stress the distinction from the more general case of complex numbers. The name originated from the belief that imaginary numbers, as well as complex numbers, do not represent directly observable quantities in nature. While this point of view is now mostly abandoned, the original nomenclature still exists. + The system is not closed under the operation of extraction of the square root either; for example, V/ is neither real nor (pure) imaginary. 44 22 BASIC ALGEBRA AND GEOMETRY OF COMPLEX NUMBERS 45 2.2 BASIC ALGEBRA AND GEOMETRY OF COMPLEX NUMBERS Ifcomplex numbers are written in the usual form a ++ ih (or a + bi) then the usual algebraic operations with them are defined as follows. 1, Addition: (ay + iby) + (@2 + ib2) = (ai + a2) + i, + 2). 2. Multiplication: (ay + ib\)+ G2 + tbe) = (ayay — bibo) + iaiby + aob1) The second rule is easy to follow if we recognize that the expressions a + ib are multiplied i in the same manner as binomials, using the distributive and associ laws, and i? is replaced by — 1. Complex numbers of the form a + i0 are tacitly identified with real numbers since they obey the same algebraic rules and are generally indistinguishable from cach other.* Complex numbers of the form 0 -+ ib are thea (pure) imaginary numbers, It is customary to write simply a + #0 = a and 0+ ib = ib. Sub- traction of complex numbers can be defined as fnverse addition so that if (a1 + 181) — (@2 + iby) = x + iy, then ay + iby = (x + iy) + (a2 t ibe) from which it follows thatt x=a,—4, and y= b,— by. An alternative is to form the negative of a complex number, —(a + ib) = (~ 1a + 6) = (—1 + OMG + ih) = ~a— ib, and reduce the subtraction to addition. The rule for division can be similarly deduced by inverting the multiplication. A shortcut method is given by the following technique: atit _ (ac + bd) + chid ct ead? be ~ ad) ac+ bd, be —ad ere ape (c? + d? = 0). Itis readily seen that the divisor can be any complex number except zero (eamely the number 0 + 10, which is unique and is written simply 0). * Ina more rigorous language, “the subset of complex numbers of the form a + iO is isomorphic to the set of reat numbers under the correspondence a + i0 © a.” t ILis tacitly postulated that x1 + ayy = x2 + pa if'and only ix) = x2 and yi = ya. 46 FUNCTIONS OF A COMPLEX VARIABLE 22 Remarks 1. The addition of complex numbers obeys the same rule as the addition of vectors in plane, provided a and 4 are identified.with components of a vector. Note, however, that the multiplication of complex nunibers differs from the formation of dot and cross products of vectors. 2. The use of the symbol i and the related binomial @ + ib is conventional, but not indispensable. It is possible to define a complex number as a pair of real numbers, {@,A), obeying certain peculiar rules, e,, the multiplication.can be defined by (@1, bia, b2) = (@1a2 — bib2, arba + arab), and so on.; It should be clear that the form a + i is just # representation of a complex umber, _It is customary to represent complex numbers by points in the so-called com- plex plane, or Argand diagram (Fig. 2.1). If we denote the complex number x + iy by, 2 single symbol 2 and write z = x + iy, then to each z there cor- responds 4 point in the comptex plane with the abscissa x and the ordinate y. This idea also leads us to the irigonometric representation of a complex number: Z= 7 (cos@ + isin 8), 4 Beraxtiy where A= /x2-F yp? and tan@ = y/x. In ia this representation r is unique (positive square = root) but @ is not. A common convention isto | demand thatt = Re a —w<@Sn, Real axis along with the standard rule of quadrants, Figure 2.1 namely, @ < Oif yp < 0. The foflowing nomenclature and notation will be widely used: If z= x4 iy = (cos + isin @ then is the real part of z, is the imaginary part of z, is the modulus of z, also known as the magnitude or absolute value of z, @ is the argument of z, also called the polar angle or phase.t ‘The number x — ip is called the complex conjugate of the number z = x + iy and vice versa. We shall denote it by 2*. We can say that z and z* reptesent (on the complex plane) the reflections of each other with respect to the real axis. + Another commonly used convention is 0 < @ < 2n. $A more precise name for 8 would be the “principat value of the argument of 2” (see p. 57). See | | 22 BASIC ALGEBRA AND GEOMETRY OF COMPLEX NUMBERS = 47 Remarks 1. The quantity zz* is always a nonnegative real number equal to [2|? or to [2*[? (which are the same), 2. The quantity z + z* is always a real number, equal to 2 Re z or to 2 Re z* (which are the same). 3. The rules (21 + z2)* = 2f +23 and (c1z2)" = 242% are evident and should be remembered. Figure 2.2 Because complex numbers obey the same addition rule that applies to vectors in a plane, they can be added graphically by the parallelogram rule (Fig. 2.2a). Conversely, vectors in a plane can be represented by complex numbers. The scalar product of two such vectors can be obtained by the rule (21-22) = Re (2]z5) = Re (2123), where it is understood that z, and z2 are vectors corresponding to complex num- bers z; and z respectively. The vector product can be obtained in a similar fashion: fer X za] = Im (2322) = ~—Im G28). Exercise. Verify the validity of the above rules for scalar and vector products. In the theory of complex variables, the expression |z1 — za] is often used. According ‘to Fig, 2.2(b) this quantity (modulus of the complex number z; — zz) is equal to the distance between the points z) and zg in the complex plane. It follows that the statement [2 — zol < (which often occurs in proofs of various theorems) means geometrically that point z is within the circle of radius R drawn around the point zg as a center (i.¢., z is in the R-neighborhood of zo; see p. 16). The foltowing two inequalities are easily proved from geometrical considerations: We fan + zal S leat + [zal (A side of a triangle is less than or cqual to the sum of the other two sides.) 2 fza — zef > [zal — zall. (The difference of two sides of a triangle is tess than or equal to the third side.) \ 48 FUNCTIONS OF A COMPLEX VARIABLE 23 Remark, 1 should be emphasized that inequalities can cxist only among the moduli of complex numbers, not among the complex numbers themselves. A complex number cannot be greater or smailer than another complex number. Also, there are no positice or negative complex numbers. 23 DE MOIVRE FORMULA AND THE CALCULATION OF ROOTS While addition and subtraction of complex numbers are most casily performed in their cartesian form z = x + éy, multiplication and division are easier in trigonometric form, If z, = ry(cos #, + isin @)) and zz = ro(cos #2 + isin G2), then elementary calculation shows that 2122 = Fyro{cos (6) + G2) + isin (@; + 82) with the provision that if #; + 62 happens to be greater than x, or Jess than or equal to —z, then the amount 21 should be added or subtracted to fulfill the condition ~+ < (6; + 62) 0) to which the positive imaginary semiaxis is added. The negative imaginary semiaxis is not included. The second branch, which has no special name, maps the z-plane onto the left half-plane (Re w < 0) plus the negative imaginary semiaxis. Except for z = 0, no other point on the w-plane (image plane) is duplicated by both mappings. Also observe another important feature of the two branches. Each branch taken separately is discontinuous on the negative real semiaxis. The meaning of this is as follows: The points ee at BY where 8 is a small positive number, are very close to each other. However, their images under the principal branch mapping, namely Siler = 2D and © fy(za) = eT HH), are very far from each other. On the other hard, note that the image of zz under the mapping f(z), namely, Soleo) — ete 248/21, is very close to the point f:(z1). It appears that the continuity of mapping can be preserved if we switch branches as we cross the negative real semiaxis. To give this idea a more precise meaning we must define the concept of con- tinuous function of a complex variable. Let w = f(z) be defined in some neigh- borhood (see pp. 47 and 16) of point zp and let f(zo) = Wo. We say that f(z) is continuous at zq if* f(z) wo whenever z— Zo in the sense that given 5 > 0 {arbitrarily smafl), the inequality |/(2) — wol’< 8 holds whenever |z — zo| <€ holds, for sufficiently small ¢. It is readily shownf that if w = u(x, y) + HG, y), then the continuity of w implies the continuity of u(x, p) and u(x, y) and vice versa. * Also written as lim..1, £(2) = f(z0). } For example, see Kaplan, p. 495. 2.6 MULTIVALUED FUNCTIONS AND RIEMANN SURFACES 55 Riemann proposed an ingenious device to represent both branches by means of a single continuous mapping: Imagine two separate z-planes cut along the negative real semiaxis from “minus infinity” to zero. Imagine that the planes are superimposed on each other but retain their separate identity in the manner of two sheets of paper laid on top of each other. Now suppose that the second quadrant of the upper sheet is joined along the cut to the fourth quadrant of the lower sheet to form a continuous surface (Fig. 2.6). It is now possible to start acurve C in the third quadrant of the upper sheet, go around the origin, and cross the negative real semiaxis into the third quadrant of the lower sheet in a con- tinuous motion (remaining on the surface). The curve can be continued on the lower sheet around the origin into the second’ quadrant of the lower sheet. Two edges joined here, Lower sheet Figure 2.6 Figure 2.7 Now imagine the second quadrant of the lower sheet joined to the third quadrant of the upper sheet along the same cut (independently of the first joint and actually disregarding its existence). The curve C can then be continued onto the upper sheet and may return to the sfarting point, This process of cutting and cross- Joining two planes leads to the formation of a Riemann surface which is thought of as a single continuous surface formed of two Riemann sheets (Fig. 2.7). An important remark is now in order: The line between the second quadrant of the upper sheet and the third quadrant of the lower sheet is to be considered as distinct from the line between the second quadrant of the lower sheet and the third quadrant of the upper one. This is where the paper model fails us. According to this model the negative real semiaxis appears as the fine where all four edges of our cuts meet. However, the Riemann surface has no such property; there are two real negative semiaxes on the Riemann surface just as there are two real posi- tive semiaxes. The mapping f(z) = \/Z may help io visualize this: The principal branch maps the upper Riemann sheet (negative real semiaxis excluded) onto the region Re w > 0 of the w-plane. The line joining the second upper with the third lower quadrants is also mapped by the principal branch onto the positive imaginary semiaxis, The lower Riemann sheet (negative reat semiaxis excluded) is mapped by the second branch onto the region Re w < 0. The line joining the second fower with the third upper quadrants is mapped (by the second branch) onto the negative imaginary semiaxis. In this fashion the entire Riemann surface. 56 FUNCTIONS OF A COMPLEX VARIABLE 2.6 is mapped one-to-one onto the w-plane (z = 0 is mapped onto w = 0; this particular correspondence, strictly speaking, belongs to neither branch since the polar angle @ is not defined for z = 0). The splitting of a multivalued function into branches is arbitrary to a great extent, For instance, define the following two functions which also may be treated as branches of f(z) = V2: Jr eft for O<@ cos z tanz : cosh 2 = $(e* + e7"), sinh z = }e* — e77), ' sinh z 1 tanh z = Sogha’ cothz = Ganz’ . * If the so-called point at infinity (Section 2.14) is taken into account, then the mapping f(@) = V7 also has two branch points, . . 26 MULTIVALUED FUNCTIONS AND RIEMANN SURFACES 57 Alt these functions are periodic: sin z and cos z have a (primitive) period 2x, tan z has a (primitive) period x, e', sinh z, and cosh z have a (primitive) period 2ni. A score of familiar formulas can be established, for instance, sin (21 + 22) = sin z, cos zg +- cos 2, sin za, sin z1 — sin zp = 2cos Zit2o.. m+ ze 5-3 sin +) ete. 2 2 Also note that coshz = cos(iz) — sinhz = —/sin (iz). It is worthwhile to mention that {sin z] and |cos z| are by no means bounded by unity, for instance, Isin 2if = 3.24. The logarithmic function is defined as the inverse of exponential function. Solving e” = z = re“ for w, we obtain the general solution w= loge} i+ ne — (n = integer). This function is multivalued: Its principal branch is usually denoted by w = log z and is defined as logz = logr+i@ (-¥ <8 < xh The entire multivalued function is referred to as w = Logz = logz + inn. ‘These formulas are often written with the help of the argument of z function which is also multivalued, the principal branch being agz=@ (-w<@<7) and the entire function reading Arg z = arg z+ 2x. Thus we may write logz = log|z| + fargz, Logz = log |z] + /Argz. The functions Arg z and Log z require a Riemann surface consi many Rietnann sheets. The definition of inverse trigonometric and hyperbolic functions now easily follows. All are multivalued: Arccosz = i Log(z + Vz? — 1), x Are sinz = 3 ~ Arccosz, ing of infinitely 1 i-z Arctanz — y;Log 743° Arsinh z = Log (z + Wz? + 1), Arcosh z = Log (z + Vz — 1), Artanhz = 4 Log $2. 58 | FUNCTIONS OF A COMPLEX VARIABLE 27 Their principal branches are denoted by arc cos z, etc.* The general power fisction 2° is defined by fom aloes, For rational a(a.= p/q, where p,q are integers and q x 0) this function coincides with ¥2? (g¢ > 0) or with (V2*)-? | < 0) and has g branches. For irrational a the power function is infinitely multivatued. 2.7 ANALYTIC FUNCTIONS. CAUCHY THEOREM In this section we shall discuss the subject of calculus of functions of a complex variable. The basic concept of the continuity of a complex function has already been presented, and it is not ifficult to verify that the sum, product, and quotient {except for division by zero) of two continuous functions is continuous, A con- tinuous function of a continuous function is also continuous.t Let C be a piecewise smooth curve in a complex plane. If f(z) is continuous on C, then the complex integral d: [foe can be defined and expressed in terms of real integrals by putting f@) = ule, y) + wl, y) and de = dx + idy; this yields fayde = [ (udx —vdy) + if @dx + udy), Cc Cc Cc where the real integrals Jo (u dx — vdy) and fo (v dx + udy) are known to exist. Curve C may be open or closed but the direction of integration must be specified in cither case. The reversal of this direction results in the change of sign of the integral. Complex integrals are, therefore, reducible to curvilinear real integrals and possess the following properties: [, Wer + a@)de =f, ferds +f, ete) de Ww GC Cc Cc f kf@dz =k , fizjdz — (k = complex constant), ' @ f, fa) dz = f, fo dz + i, fea, e where C is decomposed into two curves, C, and Co. The absolute value of an integral can be estimated by the formula | I fey «| < ML, where M = max[f(z)| on C, and L is the length of C. * Another widespread notation is arc sin z = sin7! z, arsinh z = sinh~ 'z, and so on. { For example, Kaplan, p. 496. {In the sense of Ricmann; see, ¢.2., Courant, Vol. 1, p. 133, and Kaplan, p. 299. 27 ANALYTIC FUNCTIONS. CAUCHY THEOREM 59 As our next concept we shall define the derivative of a complex function: Changing z into z + Az (with complex Az), we obtain f(z + 4z) and we can write™ fz), I) = 2 fe) = jim SEAM As in the case of real functions, this limit may or may not exist. It may be empha- ~ sized that in the above formula Az may approach zero in an arbitrary fashion, that is, z + 42 may approach z along any curve or by any sequence, This rather stringent requirement implies that f(z) must indeed be “well behaved” at point z in order to be differentiable. Function f(z) is said to be analytic (regular, or holomorphic) at point z if it possesses a derivative at z and at all points of some neighborhood of z (small but finite). This additional requirement results in many desirable properties of analytic functions, such as the existence of derivatives of all orders. The theory of functions of a complex variable deals essentially with analytic functions. Mere existence of a derivative at all points of a neighborhood may be shown to imply that the derivative is continuous.t Also, it is a simple matter to verity (by the same technique as for real variables) that the derivatives of complex functions obey the usual rules: a Hors +0) — 0 4 Oe, ) Lovin) = Mt 4 Py, @) a # where w= wr) and f= #2), @) 4 nz" (4 = integer), 3) and so on. The differentials of complex functions are defined in the same way as for real functions: If w = f(z), then dw = f"(2) dz If we set f(z) = w = u(x, y) + io(x, y), then the definition of the derivative can be rewritten as Je) = tim WA Ax y+ Ay) — wx, 9) + ole + Ox y+ Ay) — 08»), Aso Ax | iAy Ay0 The limiting value on the right-hand side must be the same for the arbitrary ap- proach Az — 0. In particular, set Az = Ax (approach along the real axis); then Pea Be i®. * See p. 54 (including the footnote) for the definition of a limit. t See, e.g., Knopp, Theory of Functions, Vol. 1, p. 63. 60 FUNCTIONS OF A COMPLEX VARIABLE 27 Alternatively, set Az = Ay (approach along the imaginary axis); then ee Lo-Fi It follows that for a differentiable function w = u + iv we must have au av au oe ox ay’ aye These are the Cauchy-Riemann conditions; they follow directly from the definition of the derivative. If, further, f(z) is analytic, then f’(z) must be continuous which, in turn,* implies that the partial derivatives of w and v are continuous. The inverse theorem also holds: If u(x, y) and ofx, y) have continuous first partial derivatives satisfying Cauchy-Riemann conditions in some neighborhood : of z, then f(z) = # + 1 is analytic at z. ao Integrals of analytic functions possess some very important properties. Per- haps the most fundamental one is expressed by the Cauchy theorem: Uf f(2) és analytic in a simply connected domain D, and C is a (piecewise smooth) simple closed curve in D, then $e f@) dz = 0. Proof. Write the integral as $f a = gua — ody) + if Ode + udy). Analyticity of f(z) implies continuity of gartial derivatives of w and v and Green's theorem (p. 26) is applicable. However, then the Cauchy-Riemann conditions imply dv au fad oan =ff (34% dx dy = 0, ai fou+uay=ff 2 dx dy = 0, 8 and the theorem follows. There is a converse of the Cauchy theorem, known as the Morera theorem:t If f(z) is continuous in a domain D and if $f(2) dz = 0 for every simple closed path in D with its interior also in D, then f(2) is analytic in D. rt It is not difficult to see that the Cauchy theorem is true for multiply/connected domains provided the interior of the simple closed path C is also inside the domain * See, e.g., Kaplan, p. 510. } Knopp, Theory of Functions, Vol. 1, p. 66. 27 ANALYTIC FUNCTIONS, CAUCHY THEOREM 61 (ie, the path does not encircle a hole; see Fig. £.15), Similar extensions hold for related integra! theorems that will be quoted later, The vanishing of a contour integral (an integral around a simple closed path) is closely related to the independence of path of an integral. In fact, the considera- tions of Section 1.6 can be applicd easily to complex integrals, leading to the statement: If ¥ f(z) dz = 0 for every simple closed path, then the integral ff fevar is independent of path (between zp and z). Suppose now that the point zo is fixed. If the integral [°, f(¢} dt is inde- pendent of path, then it must represent a function of z. This function is then a primitive function of f(z) (or an indefinite integral of f(2)) as follows from the Sundamental theorem of integral calculus: If f(2) is analytic in a simply connected domain D, then the function Fe =[" Aor is also analytic in D and f(z) = (d/dz)F(2). Proof. Since f(z) is analytic, the integral is independent of path and is therefore a function of z, In the expression ay F@)=U+iV= (de — vay) + if" (ode + udy), (4 a,¥o? 0.m0 both integrals are independent of path (by Green’s theorem and the Cauchy- Riemann conditions). It also follows that* w_, wy ax” ay , wi, My x By so that and v satisfy the Cauchy-Riemann conditions as well. Therefore F(Z) is analytic, Morcover, dF _ au, av a Bt an tigg tet w= fe), and the theorem follows. Any two primitive functions must differ by a (complex) constant; this follows from the fact that/”(z) = Oimplies /(z) = const (integrate au/ax = 0, au/ay = 0, elc.). * For example, Kaplan, p. 244. 62 FUNCTIONS OF A COMPLEX VARIABLE 28 2.8 OTHER INTEGRAL THEOREMS. CAUCHY INTEGRAL FORMULA It should be emphasized that all conditions stated in the Cauchy theorem must be checked before applications. Consider, for instance, the integral = 1 = T= fie (a = const). Is this integral zero or not? Generally speaking, f(z) = 1/(z — a) is an analytic function but it fails to be analytic at one (and only one) point, namely, z= a. The function is not even defined at this point and thus cannot possess 2 derivative. . Let the curve C involved in the definition of J be a simple closed curve. Then, if the point z = a is outside the curve, the Cauchy theorem holds and J = 0. If it is inside, the Cauchy theorem cannot be applied. In fact, the integral is not equal to zero, as demonstrated by the following considerations: If Cis a circle of radius R centered at z = a, then the integral is easily evaluated by setting z=a-+ Re®, In this case dz = iRe” do and r= [ide = dvi. It is not difficult to show that the same result is true for any simple closed path C,, which encircles point z = a. Suppose that C, is entirely inside the circle C (Fig. 2.8). Then a thin channel made up of curves By and Bz can be constructed to connect the interior of C1 with the exterior of C and the Cauchy theorem can be applied to the shaded region; a domain D can be constructed so that the shaded region is within it. The integral over C, is clockwise. As the sides By and Ba of the channel are allowed to approach each other, the integrals of {@ = 1/ — @) along B, and B, will (in the limit) cancel out, leaving us with the statement ia c ¢, Se)dz + ¢ flay dz = 0. c ° cs Counterclackwise ——Clackwise Reversing tHe direction of the second integration to > counterclockwise, we obtain A Ah 5 8; an a ee 2 fod = 4, 1@d Re Figure 2.8 (with both directions counterclockwise). : If Cis entirely within C1, the proof is similar. If C and C, intersect, the proof is even simpler. Exercise. Produce a proof of the discussed statement given that C and Cj intersect at two points. | 2.8 OTHER THEOREMS. CAUCHY FORMULA 63 If the integrat J is evaluated around a closed path which is not simple, its valuc is not necessarily 2x, In cases of practical interest it will be equal to n2zi, where 1 is the number of times the path encircles the point z = a counterclockwise less the number of times it encircles the point z = a clockwise. Of course, it should be understood that the integral $f(2) dz may happen to be zero even if the Cauchy theorem does not apply. For instance, calculate the integral t= $eige where 7 is a positive integer not equal to unity and the contour is a citcle of radius Raround z = a. Using z = a + Re®, obtain +n os sf pRingt—08 gg = PEN san —0 1 This result evidently holds for any closed path encircling z = a. In both of these examples, the possibility of the point z = a being exactly on the path of integration has been avoided, and for a good reason; such integrals cannot actually be defined. Whenever this situation occurs in practical problems, the path must be deformed to avoid the troublesome point. How this is to be done depends on the nature of the problem.* Function f(z) in the Cauchy theorem must, of course, be single valued. It may be a particular branch of a multiple-valued function, but then care should be taken that this particular branch is analytic. Consider, for instance, the integral Fines VF along the nit circfe about the origin. First of all, the branch of the (double-valued) function 1/2 must be specified. Suppose it is the principal branch. Then § Vids =f eRe ao = — 45, The Cauchy theorem is not applicable because f(2) is not analytic within the circle lz] = 1. The points where it fails to be analytic are along the real axis from x= —1 to x = 0 where f(z) is not even continuous. Note also that although £(z) is continuous at z = 0, it is not analytic at that point either. Consider now the same integral ¢ Vide #4-21=1 * One such example is given in Section 12.9. 64s FUNCTIONS OF A COMPLEX VARIABLE 28 taken around the point z = —2 (Fig. 2.9). If the principal branch is involved in the integration, the Cauchy theorem is not applicable. However, split V2 into the following two branches (as on p. 56): fz = VF eh? if O<0<5, Branch A: ‘i Vea Vr eiltael if we 8S 0. Branch B: (v3 = Vrelltti if OC OSs, ~ V2 = vrei? if -x<6<¢0. Now the branch cut is along the positive real semiaxis, and each branch js analytic within (and on) the circle |z + 2| = 1 and the Cauchy theorem may be applicd. Figure 2.9 Figure 2.10 The Cauchy theorem can be generalized in more than one way, Observe that the interior of 2 simple closed path is evidently a simply connected domain with the path serving as its boundary. A (finite) collection of nonintersecting simple closed paths may form a multiply connected domain, as illustrated in Fig. 2.10: The paths Cy, C2, and C3 form the boundary of the (triply connected) domain S. With respect to this domain S, the contours Cy, Cx, and Cg are oriented in the following sense (see p. 21): The positive direction of circulation around C1, Ca, and C, is such that the domain S appears to the feft. This implies, in our example, a counterclockwise direction for C; but a clockwise direction for Cz and C3. ‘These concepts lead (o the so-called Cauchy theorem for multiply connected domains: If S is a multiply-connected domain whose boundary B consists of a finite number of simple closed paths, then (2) dz = 0, §, 40 provided f(z) is analytic in S and on B and the integration is carried out in,a positive direction over all parts of B. The proof is constructed without difficulty by using the technique of channels (p. 62). In this (and other) integral theorems the analyticity of f(z) on @ contour fas well as in its interior) is demanded. ‘This is because the Cauchy theorem requires 28 OTHER THEOREMS. CAUCHY FORMULA 65 the contour to be located within some simply connected domain D. Actually, this condition can be relaxed and the following boundary integral theorem holds: If f(2) is analytic in the interior S of a simple closed path C and f(z) is continuous on C* then $c f(z) dz = 0. The proof is based on the construction of a contour C’ which is in $ and arbitrarily close to C; the Cauchy theorem is valid for C’. The continuity of £(z) is then used to show that the integral over C’ must approach the integral over C as C’ approaches C. The Cauchy theorem can be used to deduce many other properties of integrals, the basic one being the Cauchy integral formula: If f(2) is analytic inside and on C and if the point z = a is in the interior of C, then Me 5, BG = ifla). Proof. Construct a circle C’ about z = a with an arbitrarily small radius R such that the circle is within C (Fig. 2.11). The integral over C is evidently the same as the integral over C’ (use the channel technique). Rewrite this integral: ade _ £ fa)— fla) $08 - | MaMa no$ 5 f2)— Ka) pO da + fla\axi by a previous result. The remaining integral on the right-hand side must be inde- pendent of the radius R (because other terms are independent of R). It can also be estimated that $ LO= LO 4.) < ¢ LO = LON a,c Mage = 16M ro ~ Ser "= OR 7 zoe = al where Mf — max 1f(z) — f(@[ on the circle C’. If R is chosen sufficiently small, then M can be made arbitrarily small [by continuity of f(z) at the point z = a}. Therefore the quantity ¢ $O= fe) 4) zZ—a@ can bé shown to be less than any positive number, no matter how small. This is possible only if £2) — fla) dz =0, cz a and the theorem follows. Figure 2.11 * To be precise, provided a point on C is approached from the interior. 66 FUNCTIONS OF A COMPLEX VARIABLE 29 The Cauchy integral formula reveals a remarkable property of analytic fune- tions: If the values of a function ace specified along a closed contour, then its value at an arbitrary point inside the contour is already predetermined. To emphasize this, replace a in the formula by z (variable) and denote the dummy variable by ¢: I $fod, JO) = FGF Fz Jn this formula, the integral is a Function of the (variable) parameter z and can be differentiated with respect to z. The following theorem (Leibmitz rule) follows from the corresponding, properties of real integrals:* d — § ahh 2) £$ pode f op ot (contour C.must be a simple closed curve in the usual sense, namely, it should not extend to infinity, otherwise the convergence questions must be considered). Application of the Leibnitz rule to the Cauchy integral theorem yields an inte- gral expression for the derivative: Ye) _ 1g LOK | dz Imi l 2 Repetition of this process gives us the nth derivative Pf2)_ mf MOG. dn ~ di} (— 2H 2.9 COMPLEX SEQUENCES AND SERIES No serious study of analytic functions is possible without their representation in the form of series, and we now turn our attention to consideration of this aspect of the theory. Our first task would be to consider complex sequences. ‘An infinite sequence of complex numbers {2} = (21, 22,--} is said to converge to the (complex) limit z, provided ; len — 21 <€ for sufficiently large 7; ¢ is, of course, an arbitrarily small positive number. Con- vergence of complex sequences is reducible to that of real sequences by the fol- lowing fundamental theorem: The sequence {2,} converges to z = x + iy if and oniy if Re Zp converges to x and Im z, converges to y. Proof. \f Rezm = X,— x and Im zy = yay, then : len = [Gn = 2) iG — PS ben — 31 + [rn WL < (6/2) + /2) = € * See Kaplan, p. 289. 29 COMPLEX SEQUENCES AND SERIES or for sufficiently large n. Conversely, if zn — 2] < ¢, then zy is within the circle of radius ¢ about the point z. This implies In - xl ce and [pp — yl Zk fat * See, e.g., Knopp, Theory and Application of Infinite Series, Section 10. 68 FUNCTIONS OF A COMPLEX VARTADLE 29 is a convergent sequence. Denoting S$ = lim S,, we customarily write If the sequence of partial sums does not converge, then the series is said to be divergent. It is to be emphasized that, under certain circumstances, divergent series can be given definite meanings and such series are widely used in applica- tions.* However, the theorems derived for convergent sequences should not be indiscriminately applied to divergent sequences. A series is said to be absolutely convergent if the (real) series of moduli E lal nol is a convergent series. An absolutely convergent series is convergent (the proof is trivial). In most cases onc proves convergence of a complex series by establishing that jt is absolutely convergent. The following tests are most common. Comparison test. Mf |zn| & for all n sufficiently large, and k > 1, then ¥°z, diverges. Root test. Hf Y]z,] < k < | for all n sufficiently large, then 272, converges absolutely, and if [za] 2 & > 1 for all sufficiently large, then 3'z, diverges. Prools of these theorems are similar to those for real series; they are based on the inequalities for absolute values which are also true for complex numbers. Divergence of a series can often be quickly established by the ath term test: If z, fails to converge to zero then the series }0z, diverges. If necessary, the question of convergence of a complex series can always be re- duced to that of two real series by the basic theorem: The series 127, = Fo(%n + iva) converges to S = P + iQ if and only if 2x, converges to P and Syn converges to Q. For instance, series which are convergent but not absolutely convergent can be treated in this fashion. Complex series can be addéd and subtracted provided they are convergent. They can be multiplied only if they are absolutely convergent, the product being also an absolutely convergent series; if the series are not absolutely convergent, then we are faced with the problem of arranging the product series. * For example, see Section 6.4. {See Knopp, Theory and Application of Infinite Series, Section 45. 29 COMPLEX SEQUENCES AND SERIES 69 Terms of a complex series may depend on a complex variable z. Most common serics are power serics, for instance, Ssltz24+274274.-.. In many cases such series will converge only if z is confined to a certain region. The above series converges absolutely, by ratio test, provided [z| < 1. This series diverges, by ratio test, if |z| > 1. The ratio test is inconctusive if |z| = 1, but then the xth-term test shows that the series diverges. It is seen that the above power series converges absolutely for all points inside a circle of radius R = 1 called the radius of convergence. The concept of the radius of convergence can be applied to every power series, Indeed, if a power series is convergent on a circle of some radius 7, then it is absolutely convergent everywhere inside this circle* (by comparison test). The problem js then to find the upper bound of r which is the sought radius of convergence. Exercise, Show that the series 1 — 32-4 92? ~ 2729 + Bizt — +++ has a radius of convergence equal to $, while the series L424 22? + 33 4 Aled 4 --- has a radius of convergence equal to zero; the point z = 0 is the only value for which the series converges, If the upper bound described above does not exist, then the series converges absolutely for all values of z and is said to have an infinite radius of convergence, For example, 2 2 zt (hatha tat Partial sums of a power series represent a sequence of polynomials in z. Sequences of other functions can also be considered. The sequence {f,(2)} of functions defined in a region R {z belongs to R) is said to converge to a limit func- tion f(z) in R, provided lim fue) = fl2y for each z in R. For instance, the partial sums of the series Draltr¢ 27424. aed * This statement is known as Abel's theorem. See Kaplan, p. 350. 70. FUNCTIONS OF A COMPLEX VARIABLE 29 form a sequence of functions (polynomials) = oF I= Vz, kao and this sequence converges to the function f(z) = 1/(1 — z) in the (open) region (z| < 1 because at fal z¢ 2p $ r= <3 and . zt tim y 0 for |z| < L. For this reason the function f(z) = 1/(1 — z) is said to be the sum of the above series (for |z| < 1 only!): Frad-y) ld N, holds simultaneously for all 2 in R. In plain terms: Let us suppose we desire a certain accuracy ¢ for our ap- proximation. For some particular z, the tenth function in the sequence may suffice (ten terms of a series, if we are talking about partial sums). But for another z, the tenth function may be inadequate because the speed of convergence is slower. In general, we may need to go farther and farther along the sequence as we proceed to points where the convergence gets increasingly worse. Uniform convergence sets the end to this process. The convergence can be no worse than a certain specified degree and the Nth term will guarantee a certain accuracy for the entire region. Example. The sequence of partial sums of the series m0 ‘ i is convergent for |z| <1 but it is not uniformly convergent; the convergence becomes increasingly worse as |z} > I. However, in the region-|z| < & where k <1, the convergence is uniform, It may be “bad” at z = k, but once M is 2,10 TAYLOR AND LAURENT SERTES Ww found such that 2 0 for alln > Nand z = &, then the same value of W will hold for all other 2 with bl sk. The series of functions is called uniformly convergent (in a region R) if the sequence of its partial sums is uniformly convergent (in that region). Uniform convergence of series is most commonly established by the Weierstrass M-test: The series f(z) of Functions is uniformly convergent in a region R if there exists a series of positive constants M, such that Wi@| 1/R’ = Ra, and the statement follows. Therefore we conclude thatif Ra > Xj, then the series 42 D en(z — a)" nate will converge within the annulus Ry < |z — al < Ro. It can, of course, happen that Re < Ri, in which case our series will diverge everywhere. The following theorem can now be derived: Every function f(z) analytic in an annulus Ri<|z-al< Re con be expanded in a series of positive and negative powers of (2 — a), namely & . f@= Di ene - ay. This series, known as the Laurent series, is unique for a given annulus, and the coefficients ¢c, can be obtained from fe)d m= ie 1G — ati’ where I’ is a circle of radius R such that Ry < R < Re. Figure 213° Proof. Contract the radius Re slightly and expand the radius R, slightly to-obtain an annular region, with point z inside, to which the generalized Cauchy formuta is applicable (Fig. 2.13):* fora, Vf Ma. mr—z | dyn, f—z oat The first integral can be treated as in the derivation of the Faylor series: 1 KO _ Oz Oa 1 OS ta Lp Ane erat = y C—O ae Lg Gta , tid, Foz @ — ati e- Geo Term-by-term integration is permissible by uniform convergence. * We shall occasionally use the convenient symbols g and g¢ to indicate clockwise or counterclockwise direction of integration. 2.10 TAYLOR AND LAURENT SERIES 75 The second integral is treated by expanding I/(¢ — z) in a somewhat different geometric series: Lot i ~~ = =o Poe" (Foe i= Gs aera ~~ Eye a ae which is convergent by ratio test. Then id for _ _ 1g fox Wit, Foz” TT, 2 2 1 1 m = 2 @septi HP, Hone - 0 a. Replace m by —(+ + 1) (n must be negative) and rewrite the above as al foe. F eo bg fayag Ii Fr, Foe ni Fn, & — apr” Finally, note that the integrals a §, 4, (8 =0,1,2,..) _fodg 1 (@ — at and @ = -1,-2,~3,...) may be just as well evaluated over a common citcle I’, concentric with I, and I's, and lying within the annulus Ry < R < Ry, To prove the uniqueness, assume that an expansion te S@) = a. en(z — a)" exists and is valid in the annulus Ry < jz — a] < Re. Choose an arbitrary integer k, multiply both sides of this expression by (2 — a)~*—!, and integrate around a circle T about z = a, lying within the annulus. Then fad dz . All integrals on the right-hand side will vanish except one, for which n = k, and whose value is 2x7. Therefore ¢ fe) de = aay, , @— ae which completes the proof. 76 FUNCTIONS OF A COMPLEX VARIABLE 2.10 The part of the Laurent series consisting of positive powers of (z — a) is called the regular part, it resembles the Taylor series but it should be emphasized that (he nth coefficient cannot be associated, in general, with f(a) because the Jatter may not exist. In most applications, f(z) is not analytic at z= a. The other part, consisting of negative powers, is called the principal part. Either part (or both) may terminate or be identically zero, Of course, if the principal part is identically zero, then f(z) is analytic at z = a, and the Laurent series is identical with the Taylor series. Remark. The Laurent series is unique onty for a specified annulus. In general, a func- tion f(z) may possess two or more enlirely different Laurent series about a given point, valid for different (nonoverlapping) regions. For instance, 1 _! 2 Ba. Aicyrtitete tet O |al. Write If [z| > |al, then Ja/z] < 1, and we can expand Loe E (vir T= fz fn 2.10 TAYLOR AND LAURENT SERIES 77 Therefore Ls dal > lel. This is the desired Laurent series. The function f(z) can be expanded by this method about any point z = 5: Indeed, write ee ee | $0 = a> ED TEL FTE (b #4). Then, either =-—!_ Fe 57 @— 6)" =~ GDR & Dime @= by (lz — b| < fa — dp, or = ye ey _ _ fa\= Xe (hh > Ja ~ bp. Example 2. Rational fraction decomposition 1 9-5 -@piR ER The roots of the denominator are a = i, b = 2 (simple and distinct). There- fore f(z) fails to be analytic only at z = fand z = 2 and should possess a Taylor series about z = 0 valid for |z| < 1 ({i] = 1) and two Laurent series about z = 0 valid for 1 < |z| < 2and {z| > 2. To obtain these three series, we use the identities 2 Q4iz+ l= @- 2-2 1 | | 1 £0 erHE Sym THI (; - =): Suppose that the Laurent series valid for 1 < fz| < 2 is desired. The function 1/@ ~ 2) should then be expanded in Taylor series about z — 0 (Example 1), This series is, in particular, vatid for ! < Jz] <2. The function t/(z — i) should be expanded in the Laurent series about z = 0 valid for jz| > 1 (Example 1). This series is also vatid for < Jz < 2. If these two series are subtracted, we may obtain (multiplying by 1/(2 — i) a sevies for f(z) valid for 1 < jz} <2 which is the desired Laurent series, and Example 3, Differentiation. f(z) = 1/(z ~ 192. The method applied in Example 2 fails here because of the double root of the denominator. Among the alternative methods the simplest one is, perhaps, to observe that it i A(t). zZ-tpo a@\i—z 7% ~~ FUNCTIONS OF A COMPLEX VARIABLE 2410 The seri 3 1 Scrics boetteteteo= Det (zp < fan can be diflcrentiated term by term within the circle of convergence. Therefore —! Ho Py. ¥ » Gop Tt tPF = Lat be (al < 1). Example 4. Integration. f(z) = log (1 + z) = log [1 + z] + éarg (I + 2). This is the principat branch of the {muttivalucd) logarithmic function, The branch line extends from “minus infinity” to minus one and log (1 + z) is analytic within the circle |z| = 1. We know that ‘d i BOBUt OK pgs Therefore we may expand le 2atoet 2-24 te = dCs" da a in an arbitrary fashion: 1. f(2) remains bounded, that is, [/(2)] < B for a fixed B. 2. f@) is not bounded and |f(2)| approaches infinity, namely, if(z)| > M (any M) for |z ~ af < € (some ©). 3. Neither of the two cases above; in plain terms, f(z) oscillates in a “wild” manner. Examples of these three types of singularities (at z = ) are Case l, f(z) = sin z/z, Case 2, S(2) = Vsin z, Case 3, f(z) = et? The first case turns out to be trivial because then the limit lim,_, f(z} must exist, and if the function f(z) is defined at z = a by f(a) = fim, ., f(z), then it must be regular at z = a as well. Remark. The formuta f(z) — sin z/z does not define, in a rigorous sence, the value of the function at z = 0. The extended formula sin 2/2, 2 #0, yo =| 2-0 does define the function f{2) at z = 0 (and elsewhere). To prove the above statement concerning Case 1, observe that f(z) is analytic in an annulus p < |z — a] < R within a neighborhood of z = a. By the Cauchy theorem, for any point z within this annulus (Fig. 2.14), we have fz) = LhMat, 1 gh Ma, Qi §—2 | Ini Sy ft 2 rb We shall show that the second integral must be zero for 3 all p. If this is 80, then f(z) must approach the limit (LP) which will prove two things at once: (a) lima f(2)y exists, (b) if f(a) is defined by this limit then the rede- fined function f(z} is analytic at z = a as well. Figure 2.14 80 FUNCTIONS OF A COMPLEX VARIABLE 21 To achieve this result, write ¢ — z = ($ — a) — (¢ — a) and observe that IG - 2) - @- a 2k - 4] - ft - 4 Then, for a fixed 2, 1g Mae 2xi Jy § ~ Zz > as e—al —p? ~ [zal —p ‘The integral must be independent of p because of the analyticity of the integrand. Since it is less than an arbitrary positive number (for sufficiently small p) it must be equal to zero. The proof is now complete. Because of the described property, isolated singularities of the first type are called removable singularities. In practice, if a function is defined by a formula which fails for some isolated point z = a, then the formuta is tacitly replaced by the corresponding limit. In this sense, the functions S@) = sinz/z, g@)= er", AZ) = I/z — cotz are analytic at z = 0. The second type of isolated singularity, when [f(z)| 7 oo as z— a, is called a pole. Since the singularity is isolated, there must exist a Laurent series $e fa) = Dente ~ a" valid for 0 < |z — al < R (for some R). If the principal part terminates, ie, if the Laurent series is of the form += fay~ EX ele — ay", then f(z) has a pole of order mat z = a. Conversely, if f(z) has a pole at z = 4, it must have the Laurent series (for 0 < |z — a] < R) of the above form as seen from the following argument: Consider the function gz) = I/f@)- Unless f(z) = 0 there must exist a neighborhood of z = a where f(z) has no zeros. In this neighborhood, g(z) is analytic and lg@>0 as za (because [/(z) + oz). Therefore g(z) has a zero at z = a. This zero miust be of some definite order m, namely, ee)= YA a) OX < R). 241 ZEROS AND SINGULARITIES 8l Rewrite this as ae) = @ — 0)" © KG ~ a = aH, where 6, = Bay, and 6) 0. Now Ati. 3@) ~ @ — a ¥@) Since y(z) is analytic at z = a and does not vanish there, it follows that 1/p(2) must be analytic as well and possess a Taylor series at z = a. Then f@) = lose e n= ch ~ a). fiz) ea wes a y However, this is evidently a Laurent series with a terminating principal part so that the argument is complete. Example. The function f(z) = cscz = 1/sinz has the Laurent series valid for O<|eb 0 nor approaches infinity, for an arbitrary manner of approach: For instance, if z approaches zero along the negative real semiaxis, then |f(z}| > 0; if it approaches zero along the positive real semiaxis, then |f(z)| > 90; if it approaches zero along the imaginary axis, then | (2) remains constant but arg f(z) oscillates, and so on. 82 FUNCTIONS OF A COMPLEX VARIABLE 2.41 In fact it is not difficult to show that even in an arbitrarily small neighborhood of an essential singularity, a function f(z} assumes values arbitrarily close to any desired complex number (Weierstrass-Casorati theorem}. Also, an even more explicit statement can be proved, known as Picard’s theorem. In an arbitrarily small neighborhood of an essential singularity, a function f(z) assumes infinitely many times every complex value except, perhaps, one particular value Remark. It should be emphasized that an infinite principal part in the Laurent series implies essential singularity only if the series is valid “up to the singular point”:* 1 I i fO- Gaap* aa tet does not mean that f(z) has essential singularity at z = |. The point is that the above series converges only if |z — t| > 1. [actually represents the function 1 fa) = 3242 in the annulus 1 < |z — 1] < R (for any R). This function evidently has a simple pole at z = 1. Apart from isolated singularities, complex functions can have other types of breakdown of analyticity. The most common case is that of a branch point. Con- sider, for instance, the function f(z) = /z. For every point except the origin it is possible to construct a neighborhood and find a branch of \/z that will be analytic in this neighborhood. This branch may not be the principal one or its complementary, called the second branch in Section 2.6; rather, it may be some- thing like branch A in the terminology on p. 56. For the origin, this is not possible because any neighborhood of z — 0 will contain a portion of the branch tine (whatever it may be) where the sclected branch is discontinuous. Simitar state- ments can be made about other multivalued functions and their branches. It is evident that there can be no Taylor or Laurent series valid for the region Q < jz — al'< R (for some R) about a branch point z = a, However, Laurent series valid for R, < |z — al < Rz are sometimes possible. Example VETT This series is valid for |z| > 1 and represents a branch of the function yz? — I which is analytic in this region. The branch line joins, in this case, two branch points z = +1 and z = —1 and does not extend to infinity (Fig. 2.15). Replacing * Namely, for all points in a neighborhood |z — al < e except z = @. | 2.12 THE RESIDUE THEOREM AND ITS APPLICATIONS 33 zby ( — 1), we obtain a Laurent series centered about the branch point z = |. The Jatter will converge for |z — I[ > 2. Another type of singular behavior of an analytic function occurs when it possesses an infinite number of isolated singularities converging to some fimit point. Consider, for instance, Im 1 1 IQ) = er = Say" The denominator has simple zeros whenever =1 tl ne ! Branch the zea he (a = +1,42,.. Figure 2.45 The function f(z) has simple poles at these points and the sequence of these poles converges toward the origin. The origin cannot be called an isolated singularity because every neighborhood of it contains at feast one pole (actually an infinite number of then). 2.12 THE RESIDUE THEOREM AND ITS APPLICATIONS. Let f(z) be analytic in some neighborhood of z = a except, perhaps, at z = a itself [in other words f(z) is either analytic at z = a or has an isolated singularity there]. Let C be a simple closed path lying in this neighborhood and surrounding z= a, Then the integral Res fla) = oa $ Sle) dz is independent of the choice of C and is called the residue of function at the point z=, Evidently, if f(z) is analytic at z = @ (the point z = a is then called a regular point), the residue is zero. If z = a is an isolated singularity, then the residue may or may not be zero. Examples 1. f(z) = 1/2; the residue at z = 2. f@) = 1/27; the residue at z = 0 is equal to zero. is equal to unity. According to the formula on p. 74, the residue is seen to be identical with the coefficient ¢_, in the Laurent series +o S@) = ee calz — a)", which is valid for 0 < {z — al < R {for some R). The residues of a function at its isolated singularities find their application in the evaluation of integrals, complex or real, The basis for these applications is 84 FUNCTIONS OF A COMPLEX VARIABLE 2.12 the residue theorem: if f(z) is analytic on and inside a closed contour C except for @ finite number of isolated singularities at z = 1,42, ...,4,, which ate all located inside C, then ¢, fQ) dz = Ui xy Res f(a). Cc kel ‘This theorem is proved by the technique of cutting the channels between the con- tour Cand small circles Cy, Cy, ... around each singularity (Fig. 2.16). Figure 2.16 Figure 2.17 There is a variety of practical methods for quick evatuation of residues: FIRST METHOD. From the definition, Res fla) = 54; ¢ fa) az (using a suitably chosen contour C). This method is rarely used, but may be valu- able if the primitive function of f(z) is known and has a branch point at z = a. Example. (2) = t/r, F@) = Logz. Any branch of Log z may be chosen, bal in order to preserve the relationship dF(z) &! J@ = F@) = the closed contour must be disconnected and the appropriate limiting process must be applied. For instance (Fig. 2.17), a $ 1g tin | % = tim {log (4) — tog (B)} = 2ni. cz AoA SR OZ BoA Here the principal branch was used, which possesses a discontinuity 2n/ on the negative real semiaxis. 212 ‘THE RESIDUE THEOREM AND TTS APPLICATIONS — 85. SECOND METHOD. For a simple pole at 2 = a the following formula holds: Res f(a) = lim (2 — a)f(z). The limit involved is often obtained by simple substitution or the use of well- known Jimiting values. Example.’ f(z) = tan z/z?. Then sinz 1 = lim — 0 COSZ " tan se Res f(0) = lim z=, ee THIRD METHOD. For a pole of order m at z = a the following formula holds: grt Res f(a) = me ir lim ( I@- array} Example. f(z) = e/z*. Then Res (0) = dtin {& (4) = jlime* ‘ FOURTH METHOD. A common case of a simple pole is when /(2) has the form o(z) 42) = we’ where o(a) # 0-and ¥(2) has a simple zero at z = a. In this case Res f(a) ~ £3. Note that ifz = a is a simple zero of o(2), then ¥’(@) cannot vanish. Exanple. f(@) = e/sinz. Then Res foy= | at cos 212-0 FIFTH METHOD. Expand f(z) in the Laurent series and pick out the residue. This method is valuable if f(z) is a product of functions with known Laurent series. The series for f(z) is then obtained by multiplication. In practice, the coefficient c_, can be picked out by inspection. Example. {(Q) = et/(2 + DG — Yt The residue at z — 1 is desired. In the first step, transfer the pole to the origin: z-l=0, r=e+1, Then 86 FUNCTIONS OF A COMPLEX VARIABLE 2.12 In the second step, expand e and 1/( + w): uu Pot, e attw+ aw +e eee (all w), 1 11 1 eo 2 stathiten Sb SS -$t (ot <3). In the third step, evaluate (by inspection) the coefficient with « from the product of the above two series: ifoiié 3 (BT In the fourth step, evaluate the residuc: 4 1 ff ne sn=(G-R-% Prove the validity of the third method given above. [Hint: Represent f(2) as F@ = eP@/Z- a” and use the formula from p. 66.} Exercis The residue theorem can be applied to the evaluation of a wide variety of definite integrals, real or complex. Some of the most frequently used procedures are shown in the several examples which follow. Example 1. Consider the real integral 2 1 ' =f, Toapesey pe lel). This integral can be converted into a contour integral in complex plane by selting z= e, Then = cose = 3 (2+ 2) iz 2 Zz ‘ and : I= $ _ a, cz — py — pz) where C is the unit circle in the z-plane. The integrand has two poles: at z— p and z = I/p. If |p] < 4, the pole z = p is inside the contour, while the pole z ='I/pis outside. Only the residue at z = p is necded; it is equal to t ‘ fl therefore (pl <1). 2.12 THE RESIDUE THEOREM AND ITS APPLICATIONS a7 If |p| > 1, the needed residue is at 2 — 1/p and it is equal to yielding al 1 2 Padip git oT le. Note that both results can be combined into -poy lxd., while the integral is not defined for |p] = 1. 2 This method can be used for integrals of the type fq" R(cos 6, sin 8) d6, where R(cos 8, sin 8) is a rational function of cos @ and sin 8. Example 2, Consider the real integral dx 7 dx =f) wia- ial, eae > The integral [jy dx/(x? + a) can be treated as a portion of the complex integral Fc dz/(z? + a?) evaluated over the contour C shown in Fig. 2.18. Indeed (set z = xon the real axis): +R dz f dx f az cmb a Sp Ha SoH ae Let us estimate the integral over the semicircle Ce when R is very large: Write ee 2a” 21+ at/et If |z| = R is very large, then ja?/z?| = a?/R? is small and {1 +- a7/z?| is almost equal to unity (Fig. 2.19). To be precise, observe that [1 + a?/z?| > 4 for Im =plane rmicircle Ce Im 22 il +R Re Figure 2.18 Figure 2.19 88 FUNCTIONS OF A COMPLEX VARIABLE 2.42 R > ay/7, and consequently 1 Teapey < 2 (for R > aV2). This implies l 2 5 Tere eT avV/2). Now, employ the estimate (p. 58) f dz Qn Cr? + a R < xR max <= te eral dz lim sai = 0 Rae Jeg Z? + a Observe now that the integral ¢ dz/(z? + a”) is independent of the radius K (so long as R is greater than a) because the only singularity of the integrand within C is at z = ai and, by the residue theorem, a . . lor fate = Dwi Res flai) ~ Ini z= ~ 5 (for all € such that R > a). Consequently, if we let R > oo, we have +R dz . ax . a fete tim aia iin | ta which reduces to ar : a~ L. wpe Gr Exercise. Show that ifa < 0, then $00 , dx = 7. ’ no +a a [Hint; No new calculations need be done, just some logical deductions ] The above procedure can be applied to integrals of the type f * Po) dx, : -» OG)” i where P(x) and Q(x) are polynomials in x and (a) Q(x) should have no real zeros* * If Q(x) has reat zeros, then the integra! is not defined (see p. 63), unless particular modifications are made (e.g., p- 111). 212 THE RESIDUE THEOREM AND ITS APPLICATIONS 89 and (b) the degree of Q(x) must exceed the degree of P(x) by at least 2 (otherwise the integral over the semicircle Cy may not tend lo zero). For such integrals it is true that 40 POD gy = Oni f. 0G)” = ani Res, where 374. Res is the sum of the residues of the integrand in the upper half-plane. ‘This statement is a special case of the following theorem: Uf $(2) is continuous for [2] > Ro (Some Ro) and |zf(z)| > 0 uniformly as |z| + «, then im l2l=R fe) a= Proof. az| 2nR Fran ¢l = Uf 420 8] pm rere Uniform convergence of |z/(2)} means that |2f(2)| < (for arbitrarily small whenever |z| > R (that is, independently of the manner in which z approaches infinity), Then Ms nasore| <2 and the theorem follows. The conditions of the theorem are satisfied by the function fz) = 3 (with deg @ > deg P + 2) because (a) all zeros of Q must be within some fixed circle about the origin and (b) the condition |zf(z)| < ¢ for |z| > k (for some k) can be satisfied, Remark. The method described above can be extended to certain integrals of the type ” 9) di f FO) dx, where f(x) Is an even function of x, Example 3. Consider the real integral ” cos x dx [ xe at (a> 0). Note, first of ali, that Tn 4 feosxds. © 2 Se X* Fat The replacement of x by z will not help in this case because cos z is not “well behaved” in the upper half-plane; it is not bounded. However, the function e* is bomded in the upper half-plane because e -= ee and |e] = 1 (all real x) 90 FUNCTIONS OF A COMPLEX VARIABLE 242 while Je] < 1 (all nonnegative y). For this reason, consider the complex integral i= edz +8 6 dx + et de cE a Jp PHA Scyz bat evaluated over the contour shown in Fig. 2.20. Observe that . et dz fim | ~~ nom Jen 2? + a =0 {as in Example 2). Also, J = ni Res flai) = xi so that 49g 40 40. { e“ dx cosxdx ) . sin x dx oa COS * OX = ==e7 Pram], eta! J. ere ae @>0. Since the right-hand side is real, it follows that 40 cosxdx Ime r-f, Bee are a>). Remark. The statement to. sin x dx = x2 fat 0 could have been made immediately on the grounds of symmetry. However, if it were not teue, we would have obtained the value of this integra! as well from the imaginary part of 2r? 4 Res. Example 4. Consider the real integral r= f ae Se pal ft Sin xX 2t-w X Since sin z is not “well behaved” on the upper half-plane, we shall try to evaluate thé comptex integral | fe A new problem now arises: the integrand has a pole on the real axis. Note that this pole is not caused by sin x but rather by cos x, which has been added to form the complex integral. . As before, observe that 242 THE RESIDUE THEOREM AND ITS APPLICATIONS — 91 cr eRe +R Figure 2,20 Figure 2.21 Te be able to apply the residue theorem we must avoid the pole at the origin in some ion, Let us suppose we do this by means of a semi sadius r in the upper hal/-plane, as shown in Fig. 2.21. Then we ea of a small write The reason the chosen contour C is helpful in the evaluation of our Teal integral is as follows. Note that (by continuity of sin Xx/x} ee o (in either order), we obtain * sin x ; i . if = dx — wi. x Therefore* I >> »|2 # i via ; * Of course, this integral can be evaluated by much more elementary methods. The pur- pose of the above analysis is to iflustrate the techniques of the residue calculus on a simple example rather than to obtain this particular result. . 242 THE RESIDUE THEOREM AND ITS APPLICATIONS — 93 Example 5. We shall now evaluate the integral of Example 4 by a somewhat different method. We write ” sin tf sin a taf eae bf stay Lf Hazy, ox Zo 2h. Here we treat our integral as a complex integral over a path (real axis) which is open (50 far). Since sin 2/z is continuous at z = 0, we may deform the contour as shown in Fig. 2.22 and claim that ratio [ S102 | loz rot Now set ——+Re sinz = bee? — e* z= 5" — e7¥4), The problem is to evaluate . ev = lim saz and rea tor 2 For /,, close the contour as usual (Fig. 2.23a), Show that the integral over Cr. approaches zcro, and deduce that fe [Sass on z For /2, close the contour through the fower haff-plane as shown in Fig, 2.23(b). Now [e~“| is bounded in the lower half-plane and the integral over Cp approaches zero. On the other hand, note that (a) there is a contribution from the pole at the origin and (b) the clockwise integration introduces a change in the sign. When this is taken into account, we obtain f s- de = —Ini Res f(0) = —2ni. 1 Im Tm Ce a —-_+ Re Figure 2.23 94 FUNCTIONS OF A COMPLEX VARIABLE 2.12 Combining these results, we deduce that 40 sinx 1 1 E SA de = 55h — fe) = 30+ 28) = & in conformity with the previous result. Remark |. The integral over the semicircle C, in Example 4 has been shown to yield the value —zi in the limit ry > 0. This is just one-half the value oblained by integration over the fult circle. One may prove a general theorem to this effect: Let f(z) be analytic at z= a, Consider the integral [Woe nna taken from z1 = @ + re to z2 = a+ re along the circle \z — al =r (Fig, 2.24). Then lim 1, = aéffa) ro Figure 2.24 where a = 02 — 6; + 2nm (choose nso that la] < 2x). Exercise. Prove this theorem using @ technique similar to that in Example 4. Remark 2. The statement that the integral of eit/z over Cr in Example 4 vanishes in the limit R —+ © is also a special case of a more general result, known as Jordan’s lerama: if f(2) concerges uniformly to zero whenever 2 approaches infrity, then tim f ftzye™ dr Ram ICR 0 where ) is any positive nwnber and Cx is the upper half of the circle The term “uniform convergence” as z—» © means that |/{z)] < « whenever |z| > M (for some M) no matter what the phase of z is, Example 6, The integral reads +7 or e rf cosh xx (ow on. St follows that J; is bounded by, say 2e'""+””, and this approaches zero since lal < x. A similar argument holds for 2. Collecting our results, we obtain 12 2m =IntlatAtia= te so that = 1 —— = a, ~ cos (ajay ~ 3 96 FUNCTIONS OF A COMPLEX VARIABLE 212 Cfixample TD) This example deals with the extension of previous techniques to contours involving branch lines of multivalued functions. Consider the real integral “es ref Me @ 0 and x —> 00. In the field of complex numbers the general power function z*~* (p. 58) implies ZIT) = Ge MEogs o [zfpmteile Ares os [af bella Met ake? The branch which reduces to the real function x7~* (for x > 0) is the principal branch, with k = 0. It is convenient to work with the branch for which O< Argz < 2r. This branch coincides with the principal branch for Im z > O and with the branch k = 1 for lmz < 0. Then our real integral £ coincides* with the integral of this branch along the upper edge of the branch cut (see Fig. 2.26). Close the contour as shown in the figure. The integral over the lower edge of the branch cut is then ° ‘ - [etree bee _ rons f wt dx . [ rer ¥= ne lo T+x Therefore, ede golds co, V2 Seg 12 Ii Res f(—1) = [LE — oe Now, for [z} = R, roa ce i +z | ~ 2m R79 (as Ro) and for [el » 7, [ od dz! cp bbe The residue at z = —1 is e@—*; therefore vw Dart = 2x7 30 {as r > 0). jeter * In the limit ¢—+ 0,7 = 0, and R-> =, of course. 2.13 CONFORMAL MAPPING BY ANALYTIC FUNCTIONS 97 Figure 2.27 Remark. Itis also possible to operate with the standard Principal branch —r < Argz 0) onto the open region consisting of the w-plane with the real positive semiaxis removed. This is usually referred to as the ““w-plane with acut along the positive real semiaxis.”” Now consider a point 7 = zo in the z-plane and a smooth curve C passing through it. Let f(z) be analytic at 7 = Zo and let f(z) map point zp onto wo and the curve C onto curve I’ passing through it (Fig, 2.30). Also consider a point z, on C, close to Zo, and its image w, on I. Denote 2— 29 = Az, W1 — Wo = Aw. By definition of the derivative, . AW 1, sim az = f'(20)- Therefore . jan] lim = = Mf'o)l- itm Taal \F'@o)] Geometrically, the left-hand side can be interpreted as the magnification of an (infinitesimal) arc of curve C as the curve C is transformed into the curve Pr. This magnification is the same for all curves passing through zo. Now consider the angles 9 and $9 which the curves C and T make with the real axis, We have oo — % = lim [arg Aw — arg Az] = fim (are 42) = arg f(zo) — & 280 a9 Geometrically, this means that the curve C is rotated through an angle a when it is transformed into T. This angle of rotation is the same for all curves passing through zo. Observe, however, that these two geometrical statements lose their meaning at the points zo where f"(zo} = 0. - ‘The properties described above are usually referred to as the conformal mapping properties. Customarily the term “conformal mapping” is defined by: 213 CONFORMAL MAPPING BY ANALYTIC FUNCTIONS — {0 Im Im ¢ Yr 2 " % o> plane oy lane Re Re Figure 2.30 a) Invariance of angles: An angle at zo formed by two curves remains unchanged {in magnitude and orientation). b) Invariance of infinitesimal circles: An infinitesimal circle around Zo fetains its shape; it differs from a circle by higher order infinitesimals than its radius, Our analysis leads then to the fotlowing theorem: if a function f(z) is analytic at z9 and the derivative f’(2o} does not vanish, then the mapping z— f(z) is con- formal at 25. Remarks 1, The statement /’(z9} + 0 guarantees the existence of the inverse mapping because if w= fQz} and z = g(w), then g’(wo) = 1/f'(z9) Provided f’(zo) = 0. 2. The mapping f(2) — z*, which reflects the zplane in the real axis, preserves the angles and maps circles into circles but it is not considered conformal because the orienta- tion of the angles is reversed (Fig. 2.31). im Im Ce ‘ G L Ty Ty re {——____“li_p, Conformal mapping Im o Ign . Cy Re re Re r iy Figure 2.31 Mapping flz)==* 102. FUNCTIONS OF A COMPLEX VARIABLE 244 2.14 COMPLEX SPHERE AND POINT AT INFINITY Many concepts of the theory of complex variables are greatly simplified through the introduction of the so-called point at infinity. This is done with the help of the stereographic projection between the complex plane and the complex sphere which is defined as follows: Construct a sphere of radius R (for convenience, R may be taken as 4) such that the complex plane is tangential to it at the origin, as shown in Fig. 2.32. The point P on the sphere opposite to the origin (called the north pole for convenience) is used as the “eye” of stereographic projection. Lines through P are drawn which inter- sect both the sphere and the plane per- ' mitting a mapping of point 2 on the plane = Image of onto the point ¢ on the sphere (Fig. 2.32). “BAY NE In this fashion the entire complex plane is mapped onto the complex sphere (or Riemann sphere). Curves and regions in the z-plane are mapped onto curves and regions on the f-sphere. Note that the point P itself has no counterpart on the z-plane. Nevertheless, it has been found convenient to adjoin to the z-plane an extra point, known as the point at infinity, in such a way that a curve passing through P on the ¢-sphere is defined as approaching the point at in- Figure 2.32 Jinity in the z-planc. ‘The concept of the point at infinity is very useful, particularly in the analysis of mappings. The following statements of properties can be verified without much difficulty. 1. Circtes in plane are mapped onto circles on the sphere which do sot pass through P. North pole P Image of reat axis The origin 0 N . Straight lines in plane are mapped onto circles on the sphere which do pass through P. 3. Maps of intersecting straight lines have two common points on the f-sphere, one of which is P. 4. Maps of parallel straight lines have only the point P in common and they have a common tangent at P. 5. The exterior of a circle Jz| = R with R >> | is mappcd onto the interior of « small spherical cap around point P. As R-> w the cap “shrinks to, ” These and similar properties give rise to a series of definitions applied to the so-called extended complex plane, ie., a complex plane to which the point at infinity is adjoined. 244 COMPLEX SPHERE AND POINT AT INFINITY 103 Examples a) The mapping w = I/z maps zp = 0 onto wo = a (point at infinity adjoined to w-plane) and vice versa, The rigorous meaning of this is, of course, that if a sequence of points in the z-plane converges to 2) = 0, then a correspond- ing sequence of points on the w-sphere converges to its north pole. b) The region |z| > R is a neighborhood of the point al infinity, The importance of the complex sphere is greatly enhanced by the fact that if (wo curves intersect in the z-plane at an angle 7, then their images on the sphere will intersect at the same angle. In fact, the stereographic projection is conformal. (The proof is not difficult, but will be omitted here.) This permits the definition of the angle at infinity which two curves make if they recede to infinity in the z-plane. This angle is defined fo be the angle that their images on the sphere make at the point P. Now the following theorem can be stated. ‘Theorem. The mapping w = \/2 is conformal ai the origin z = 0. Observe that the function f(z) = 1/z is not defined at z = 0 but the mapping w = 1/z is defined (by the subterfuge of the sphere). The conformality does not follow from the analyticity, but rather from the relationships on the complex sphere. Corollary. The mapping w = 1/2 is conformal at infinity (despite the fact that f"(z) approaches zero as z recedes to infinity). The concept of the point at infinity is closely interwoven with the study of singularities of analytic functions. The very notion of analyticity can be extended to the point at infinity by the following device: A function f(z) is defined to be analytic at infinity if the function 8@) = f(l/2) is analytic at z = 0. Moreover, it is possible to introduce the concept of a pole at infinity, branch point at infinity, etc., through the corresponding behavior of g(z) at the origin, In this connection, the function f(z) = ¢*, which has no zeros and no singularities in the entire complex plane, turns out to possess an essential singularity at infinity. Other functions which have no singularities (e.g., all polynomials in z) are also found to have a breakdown of analyticity at infinity. Exercise. Show that a polynomial P(z) of degree # has a pole of mth order at infinity. In fact, a survey of familiar functions reveals the fact that functions which are analytic at infinity possess at Icast one singularity elsewhere, i-e., for some finite value of z, The natural conjecture is that there may not be a “perfectly analytic” function. This problem has actually been resolved and is embodied in the Liouville theorem. The only function f(z) which is analytic in the entire com- plex plane and the point at infinity is the constant function f(z) = const. 104 FUNCTIONS OF A COMPLEX VARIABLE 215 Jn conclusion, it may be mentioned that in some texts the term complex plane is tacitly assumed to mean the extended complex plane, with the point at infinity included. Certain theorems may then be more conveniently stated. However, one should never forget that while there is a point at infinity, there is still no such thing as a complex number “infinity,” in the sense that it shoutd possess the algebraic properties shared by other complex numbers. 2.15 INTEGRAL REPRESENTATIONS It is a very common occurrence that certain functions are represented and even defined by integrals, with constant or variable limits. For instance, consider the real function, the derivative of which is equal to sin. x/x. It is known that the indefinite integral of sin x/x is not expressible in terms of elementary functions* in finite form. Therefore, there is no other alternative but to define such functions by means of an integral and it is customary to write Si@)= [ a as, where Si (x) is a new function, called the sine-integral. It is, of course, but one of many primitive functions of sin x/x (ie., the one which vanishes at the origin) the others being obtainable by adding an arbitrary constant to Si (x). If we want to extend the definition of Si (x) to complex variables, it can be done in a trivial way. We write Si@) = [% sing ay, where { is a complex variable and the integral is now a curvilinear integral in the complex {-plane over some path connecting the points ¢ — 0 and ¢ = It does not matter which particular path is chosen since the function f(g} = sin ¢/f is analytic for all (finite) vatuest of ¢ and the integral is independent of path. A different situation is encountered in the so-called cosine integral, defined by . 40 , cw = -f om at. This function is easily seen to be a primitive function of cos x/x. The choice of limits is conventional but the main feature is that the integral diverges at x = 0. While it is still possible to extend the definition to complex variables by writing ‘ . ‘ cw--| eet ar, ‘ * That is, algebraic, exponential, trigonometric, and their inverse functions. + See the discussion on pp. 79-80. 245 INTEGRAL REPRESENTATIONS —105 the question arises as to the specification of the path of integration. Presumably, the upper implies an asymptotic approach to the real axis.* Three such paths are iHustrated in Fig. 2.33. The integrals taken along C, and C, are easily shown to be identical, but the integral along C, will differ by 2i times the residue of the integrand at ¢ = 6. This means essentially that Ci (z) is a multivalued function, the integrals along different paths yielding different branches. Two of these branches are characterized by the paths Cy and C; (or other paths equivalent to them). Other branches can be obtained by circling the origin m times before proceeding to “plus infinity.” Im Gi Poe, Re Figure 2.33 Remark, The above analysis reveals that Ci (x) for rea! negative x should be defined by one of the branches of complex function Ci (z) because the réal integral * cost [ 7% is divergent for x < 0 due to the behavior of the integrand at £ = 0. It has been found, however, that this divergent integral (like many similar others) can be given a defini! meaning by the prescription ~* cos § * cos § tin | f hat oat al ead known as the Cauchy principal value (or simply the principal value) and is often denoted by ‘. Because-of this result, it is customary to define Ci (x) for x < 0 by * cost Ci) = -f Se <0), Principal values of integrals occur very often in physical applications; the complex yersion of this definition} will be given shortly. * As usual, we require Ci (z) to reduce to Ci (x) as z becomes real and positive. The method of approaching infinity cannot be modified because cos z has essential singularity at infinity (see p. 109). t The integral — f° (cos £/£) df is stilt a branch of Ci(z), although not one of those described previously. {06 FUNCTIONS OF A COMPLEX VARIABLE 2.15 A second type of representation of functions by integrals occurs when the limits of integration are fixed but the integrand depends on a parameter. A well- known example of this kind is the integral f oO de = % you fo occurring in the kinetic theory of gascs and many other branches of physics.* It may be viewed as an integral representation of the function (4) = G/x/2)A7¥ 2. This point of view is, as a matter of fact, utilized in practice when the function IQ) is differentiatedt to yield other important integrals, like iQ) VE are -f te de a a —~}, 2 dy PIR) _ IVR. 512 f Aj-is? Gee EEN mf te dr, ete, In this particular example the function /(A) represented by the integral is a familiar one, but the same idea can be used to generate a variety of new functions. Con- sider, for instance, the integral Se) = fP rena (a = 1,2,3,..9, fo which is equal to n! as is not difficult to prove. ‘The above integral representation of n! suggests the extension of the notion of factorial, We may define the factorial function 11(x) by the integral I) = f "Fe dt, where x is not necessarily a positive integer. The integral converges at infinity for all values of x, but we must demand that x > — I for the integral to converge at ¢= 0, Consequently, the function M(x) is defined for all {real) x > —1 by the above integral representation. Since 11(1) = n!, we have H+ = @+)0@ @= 1423... ‘This formula holds for all values of x(x > —1). Indeed, integrating by parts, we obtain f lent dy = eet +f" e+ Diet de. ; 0 oth * For the evaluation of this integral, sce Kaplan, pr. 218. { Differentiation of improper integrals should be justified by appropriate theorems. See eg., Kaplan, p. 379. 245 INTEGRAL REPRESENTATIONS 107 Since eH lim & teow lim Hef = 0 (x> 1), io 0 (allx), it follows that [ortteas = Dl reat 0 0 Nee + D= e+ MG) & > -1). In particular, 11) = 1. This is the actual reason for the commonly used convention O=1 The recursion formula T(x + 1) = (x + IMC) permits calculation of I(x) for any x (x > —1) provided a table of II(x) is compiled for the interval 0 < x < |. Moreover, it permits an extension of the definition of (x) into the region x < ~1. Example. Calculate 1(4), 1I(~ 4), H(—3). The first valuc can be obtained directly from the definition ng) = [rae dt 0 by the substitution ¢ = x7. Then 14) = af xe? dx Ye Me by a previous result. The value of II(— 3) can also be obtained directly in the same manner, But it can be, alternatively, obtained from the recursion formula, If we setx = —4, then HG) = Hi(~-4): s(— 4) = ve. The valué of M(—%) cannot be obtained directly (the integral diverges) but it may be defined via the recursion formula, If we set x = —¥, then M™1(—4) = (— HU(—9), therefore yielding I -§) = —2Vx. An important feature of [I(x) is that it approaches infinity as x approaches — 1 from the right. The conclusion that I(x) has infinite discontinuities at x equal to. a negative integer follows from this fact and the recursion formula. The graph of I(x) is sketched in Fig. 2.34. 108 FUNCTIONS OF A COMPLEX VARIABLE 215 While the extension of the concept of the factorial is widely used in applied mathematics, it is not usually accomplished by means of the factorial function TI(x) but rather by means of the related ganna function defined by Tx) = MG - 1). Evidently I(x) satisfies the relations Te + 1) = oF) (real 9), Tx) = & — YE x = positive integer), and has the integral representation TQ) =f fed (x > 0). The behavior of P(x) a a function of x is easily obtained from Fig. 2.34 just by shifting the origin to the point x — —1,y — 0. IG} Figure 2.34 Integral representations of the type given for factorial and gamma functions can be extended to complex integrals as well. A complex function g(2) may be defined by a definite integral ata) =f” fle, s) ds, 34 Along © where the endpoints s, and s, as well as the path C between them is prescribed. In many cases the points s, and s2 coincide and the path is a closed contdur. From the theory of analytic functions, it follows that the path of integration can be deformed within certain limits (as much as the Cauchy theorem and its consequences permit this), 2.15 INTEGRAL REPRESENTATIONS 109 Im Im L Cc — Re 4.» Re fay &) Figure 2.38 Example 1, Function f(x, ¢) is analytic in the ¢-plane except for isolated singu- larities at s = 0,5 = l,ands = —1. Define a) =f fond, where € is the ‘contour shown on Fig. 2.35{a) and both limits are at infinity. Is it correct to deform the contour into another contour F shown on Fig. 2.35(b)? Although the new contour does not enclose any new singularities, it is to be remembered that the Cauchy theorem has been derived for finite contours only. This case calls for caution. Im a fs) Figure 2.36 Consider the contour [’ (Fig. 2.362) with the finite endpoints, It can be deformed as shown on Fig. 2.36(b). If the contribution from the are 5/54 of the circle of radius R approaches zero as R— <2, then the equivalence between C and Lr holds. Otherwise, it does not. Example'2.. Suppose we desire to extend the integral representation of the gamma function to complex variables. The integrand {2,0 = tet isa multivalued function, where both z and fare treated as complex, with a branch point at f= 0. The branches are given explicitly by flay 8) = e8 TP tant ot Let us select the branch of Logs for which 0 < Args < 2x, with a branch line along the real positive semiaxis and construct a contour like the one shown in 110 FUNCTIONS OF A COMPLEX VARIARLE 2.15 Circle C, of | radius +Re >Re Branch cut Branch cut fa) (by Figure 2.37 Fig. 2.37(a). It is not hard to see that the contribution to the integrat i. ve" dt from the (incomplete) circle C, vanishes in the limit when r > 0 and the integral over the upper edge of the branch cut reduces to —I'(z) if z is real* The integral over the lower edge reduces to [(z)e*—77* or, what is the same, to D(z)e"2*4, Consequently, the entire integral is equal to (e??*! — 1)P(z) for real z. It is now evident that we may deform the path C’ to path C (Fig. 2.37b) without affecting the value of the integral in any way. Moreover, the integral is well defined not only for real but also for complex values of z. This permits us to define the com- plex function T(z) by means of the integral representation ! f Poe de Cc with the path C and the branch of the integrand prescribed above. f@= Remark. This formula seems to suggest that P'(2) has poles for z equal to an integer. This is actually true only if z is a nonpositive integer: If z = J,2,3,..., then the inte- grand becomes a single-valued function analytic at ¢ = 0 and on the seal f-axis, This causes (he integral to vanish; the gamma function is then defined by a limiting process which does indeed yield (n — 1)! A great number of complex integral representations used in physics have the peculiar feature that the path of integration appears to be passing through the poles of the integrand. We shall briefly discuss the associated questions. Example 3. Let the function} f(x, k) be analytic in the k-plane except for the simple poles atk = ky and k = —ko {Ky is a real positive number). Consider x= ix, k) dk, 2) =f fo. where C is the real axis (traversed from left to right). i * The technique is very similar to that employed in Example 7 of Section 2.12. Note, however, that here we do not close the contour C” via the targe circle Ce (why?). t For example, f(x, k) — e*#/(4§ — k?) (a similar onc is used in Section 12.9). 2.15 INTEGRAL REPRESENTATIONS vt Im Im k-plane A-pline ake +k, Re Re —hy tho Figure 2.38 As it stands, the function g(x) is not properly defined because the path of integration traverses (two) singularities. However, we can give g(x) a definite meaning by deforming the path so that the poles are avoided. For a given pole, this can be done in two obvious ways: 1, avoid the pole counterclockwise, 2. avoid the pole clockwise, by “stopping the integration” a short distance in front of the pole and going, around it in a semicircle of “small” radius. In this case, which is most common in physics, this implies “going below” or “going above” the pole (Fig. 2.38). In cither of the two interpretations the value of radius ¥ does not matter so long as f(x, k) is analytic for 0 < |k + ko| 0). The path going above the pole will have a contribution —ri Res f(x, —ky) from its semicircle, Remark. In physical literature, the following notation is widely used for these two cases: Let Alay hy = £04 IVE A ko) Then the integral avoiding the pole from below is written as 40 kG, AY . = = f= ik b, a1) [ Ire re — thisreab, implying that, instead of the path going below the pole, the pole itself is displaced upward by an infinitesimal amount «. Similarly, the integral avoiding the pole from above is written as. ae We, k . a(x) -[- ii Bred. If the ideas behind this notation are clearly understood, no confusion should aeise. 112 FUNCTIONS OF A COMPLEX VARIABLE 245 Remarks 1, In the example considered, there is another simple pole at k = -+ko. It may be treated in the same way as the pole 4 = —ko but independently of it. In other words, paths which avoid one pole from below and the other one from above are widely used in practice. 2. Ifthe poles are not simple, the above analysis does not apply and the problem becomes more complicated. 3. The above considerations are easily generatized for the case of a simple pole not on the real axis, provided the path has a tangent at the pole. In problems arising from physical considerations the choice between avoiding a pole clockwise or counterclockwise (or “from above” and “from below”) is dictated by additioal requirements or assumptions, not contained in the format derivation of the integral representation (see, e.g., Section 12.9). Sometimes the physical considerations demand that the residue at the pole should not contribute to the integral at all. In this case, the integral of Example 3 may be defined to mean (as far as the pole at k = —ky is concerned) BoC) = fim { (a Sl, ky dk + . Fix, bah where « > 0; note that € must be the same in both integrals. This definition is the complex version of the concept of the Cauchy principal value. It is not difficult to verify that the above definition implies g(x) = HeiG) + gad] As our last example of integral representations we shall consider an important case of a rather simple real function expressed through a complex integral. Example 4. The step function, Let t be real and let s be complex. Consider the integral in the s-plane im 5 Mah Dei (¥> 0,4 > 0) taken along the vertical line x — ¥ as shown in Fig. 2.39(a). This particular path, often called the Bromwich contour is important in the theory of Laplace transforms (see Section 5.9). The above integral, if convergent, defines a real function (1) which we will now proceed to evaluate. Draw a circle about the origin with a large radius R and close the contour as shown in Fig. 2.39(b). First of all, integrate by parts:* at wre Mog e Le! 1 fe i fds= =) 4 - is. INS S tly thy s* Along Cp Along Cy - * Compare with p. 91. 2.15 INTEGRAL REPRESENTATIONS =— 113 Im Im Bromwich contour N y _ Cx | Re Re >Re R aby y M (ay by Figure 2.39 The integrated term te Lemal ef eit TF Ti ie tos jy ih £ will approach zero as R — on because {e™'/i| is independent of R while ihe + IY + ik In the remaining integral, estimate a 1 fe I! [ Sal < The quantity Je'"] will be bounded if ¢ < 0 (since x > 0 on Cp) and then this term will also approach zero as R— . Since the integrand e!/s is analytic within the whole closed contour, the contour integral vanishes; this implies 1h | lim 5 hase 28 Jyh In other words, our function f(¢) is zero for all negative values of 1. For ¢ > 0, the contour of Fig. 2.39(b) is not very useful since the contribution from Ce is difficult to evaluate. However, since it is the magnitude of e'* which seems to govern the estimate of the integral, it is reasonable to close the contour “to the left” for ¢ > O as in Fig. 2.39(c). As before, integrate by parts and argue that the integrated part yields no contribution (in the limit), being unaffected by the sign of 4, The remaining integral, ” 1 e tf, sts, Alan Cy L det I] Re aR (x= Res). 0 (ifr <0). 14 FUNCTIONS OF A COMPLEX VARIABLE 2.15 is estimated as follows: For the semicircle from N’ to M’ we have ef st precisely because now ¢ > 0 while Re s < 0. The length of the semicircle is +R so that the contribution vanishes. For the portions from N to N’ and from M’ to M we can at least say that ce < 2. R? and there is no contribution either. Within the contour, however, there is now a pole at the origin with the residue unity. Therefore THI ty fim tf zeal (ift > 0). In other words, f(1) is unity for aff positive t. Ht remains to evaluate f(t) for t = 0. This may be done without recourse to residues: 74ib 1 ds_ 1 Yih £0) = Jim 5 Jae 7 mt NB ih ni ales CD= Observe that it is essential that in order to obtain the value 4, the points N and M must remain symmetrical with respect to the real axis as they recede to infinity. This is somewhat analogous to the process of defining the principal value of an integral. THI og [eee yaie S ‘The function f(t) defined by (a convenient notation for the limits) is called the step function (or the unit function) and is denoted by S(#): 0 @ <9, ‘ St) = 44 @= 0), 1 @>0). Remark. The term “step function” is more descriptive and avoids confusion with the function f(t) = 1, which is unity everywhere, In many books the value S(0) = 4 is not specified and the following definition is used: 9 <0), ' Om Geo. Except in the notation used in proofs of certain theorems, this distinction is usually irrelevant in applied mathematics. 2.15 INTEGRAL REPRESENTATIONS M5 The function S(¢ — a) which is de- Sut—a) fined by 0 (0), Figure 2.40 tepresents a step function which has been shifted by an amount a. It is shifted to the right if @ > 0 as shown in Fig. 2.40, This function is called a step function in a general sense and is widely used in physical applications. It is evident that 1s ae ac) St a= a5] eds Functions given by complex integral representations are often subject to differ- entiation and integration knder the integral sign. The validity of such procedures is elucidated by the following two theorems which we quote without proofs. Theorem on integration. Let Fr and C be finite paths (open or closed) and let A(z) and f(z, 5) be analytic, with respect to z and s, on I’ and C. Then f ° hz) de f * fiz s)ds = f " ds {' ? Me2yfte, 5) de. Along t Along © Along? Along F Remark. I€T or C (or both) involve the point at infinity, then the uniform convergence of A(z)f(2, 5) is usually required for the generalization of the theorem. ‘Theorent on differentiation (Weierstrass). Let az) = f * flz,s)as. Along € If Cis a finite path in the s-plane and f(z, s) is analytic with respect to z in some closed region D (in the z-plane) for all s on C, then a) g(z) is analytic in the interior af D, b) dgtey/de = fr @f, 89/92) ds (for all z in the interior of D). ome Again, if the path C involves the point at infinity, then uniform convergence of the differentiated integral is usually required for an extension of this theorem. For instance, the. integral 1 ft te SW) = 5-5 s> ds 2wityiiw £ should not be differentiated under the integral sign. Indeed, the “differentiated integral” 1/2ni [742 c!* ds does not converge at all, much less uniformly, for any value of ¢. Note that this is true despite the fact that the derivalive dS(}/d¢ exists for all except ¢ = 0. 116 FUNCTIONS OF A COMPLEX VARIABLE BIBLIOGRAPHY Buesereacn, L., Conformal Mapping. New York: Chelsea Publishing Co., 1953. Cnurcnus, R. V., Complex Variables and Applications. New York: McGraw-Hill Book Co., 1960. Dertuan, J. W., Applied Complex Variables, New York: Macmillan Co., 1965. Knopp, K., Theory and Applications of Infinite Series, Glasgow : Blackie & Son Ltd., 1963. Knorr, K., Theory of Functions. New York: Dover Publications Inc., 1945. Koper, H., Dictionary of Conformal Representations. New York: Dover Publications Inc., 1952. Netarr, Z., Conformal Mapping. New York: McGraw-Hill Book Co., 1952, Trrcuimansn, E. C., The Theory of Functions. New York: Oxford University Press, 1950. PROBLEMS . Show that Re z1 Re z2 = § Re (z1z2) + $ Re (z1z9). Frove that if (i — 1) = —2*(i + 1), then arg z is either ¥/4 of — 30/4. . If a and 4 are (wo complex numbers and ¢ is a real parameter, then the expression z= a + 1b — a) represents a curve in complex plane. Set z = x -+ iy and de- termine the parametric equations x = x(1) and y = y(r) of this curve. What kind of curve is it? Show that the equation z = Ae‘ + Be-*, where A and B are complex constants, and f is a reat parameter, represents an ellipse. Describe this ellipse (its semiaxes, center, orientation, etc.) in terms of A and B. . Show that the transformation ey & w z+a=f (a= complex constant) represents a translation in the complex plane. Show that the equation Gato t1taet represents a circle in the complex plane. [Hinti: Setz + 1 — i= 0.1 6. Show that the equations afe--k+1) = represent, respectively, a hyperbola and a parabola in the complex plane. Can these conclusions be drawn just by inspection of the equations without any algebraic work? [Hint: Recall certain fundamental properties of hyperbola and parabola.] 7. Prove the inequalities on p. 47 by purely algebraic methods (without recourse to the Argand diagram). (Suggestion: First prove the inequalities a) i+ 2h <1+ fl, by fr - 22 W— ie} and then generalize them by setting z = 22/21. (Whal ifz1 = OD} + b) Re(l— = 10. i. 12. 13. 14. 15, 16. 7 PROBLEMS 117 . Use DeMoivre formula to show that sin 44 = 4cos? @ sin @ — 4 cos 6 sin? 9, What is the corresponding expression for cos 40? Express the following complex numbers in trigonometric form: ajz=Vi-1, b) z= (-1-"*, Show that # = e~(/2+28—) (4 = integer). {#int; Calculate Log i} Using the definitions of complex functions sin z and sinh z, show that a) sin (zy + 72) = sin z; coszz 4 cos z:sinz>, b) sinh (21 + z2) = sinh 2; cosh zo + cosh z1 sink zo. ‘What are the corresponding relations for cos z and cosh 2? Show that the complex function sin z can vanish only on the reat axis and, in par- ticular, at points x = nr (1 = integer). At what points does the function sinh z vanish? Along which curves in the complex plane do the functions sinh z and cosh z assume. real values? Show that all solutions of the equation sinz = 1000 are given, approximately, by the formula z= (tt dr i+ 7.601, where n is an even integer. [Hint: Decompose sin z into real and imaginary parts and solve the resulting equations.} Wer = xt Jy, show that +e sinh 2x -f isin 2y_ tanh y anh 2 = Cosh 2x 4 cos 2y There is no difficulty in cafculating the sums x ” a) do e™ b) De (x = real) 0 Fon) because both of them represent a geometric progression. Apply Euler's formula to deduce that x . sin }(N +1 N: 3 cos me = HHH + Dx 4, Ne = real). ks sin dx z= Teall What is the corresponding sum involving sin 2x? Apply the idea of the preceding problem (o calculate the infinite sums (consult Section 2.9) 2X 2” cos ax, a” sin nx, (a = real const). a a=0 For what values of a will the results be valid? 118 FUNCTIONS OF A COMPLEX VARIABLE 18. Figure 2.41 represents a portion of an AC-circuit 4 L with the potential difference (voltage) between points A and B given by off} = Va — Va = Vocoswt. | € Consult a textbook of your choice and show that the RB current i() (as shown) must obey the differential = 8 equation Figure 2.41 FIO pA , WO _ ae wt ater ae Show how the method of complex exponentials can be used to find a complex solution i() which varies harmonically in time. Perform the transition to real variables, show that iQ) = Ip cos Gol + 6), and evaluate Pp and ¢ in terms of £, R, C, and Vo. 19. Do the function§ x + +9” uk, = Tee? OO = —— at + C+ +3 satisfy the Cauchy-Riemann conditions? Can and v serve as real and imaginary parts of some analytic function of r = x + iy or not? 20. Consider the function S@) = Gy? — x3) + Gay? ~3yx). a) Are the Cauchy-Riemann condilions satisfied at each point of the real axis? b) Are the partial derivatives ae’ ay ax ay continuous at each point of the real axis? ©) Js the function /(2) analytic at each point of the real axis? 21, Show that the function f(2) = x? + iy is not analytic for any z. Consequently, the integral f** (2) dz is expected to depend on the path. Illustrate this by cal- culating the integral along three smooth paths of your choice connecting the points = Oandz2=2+i. * a ' t-[ 5 22. the integral independent of path? If so, state clearly under what conditions. How would you calculate J for any given path C in the most convenient way? 23, 24. 25, 2 z 28. s = PROBLEMS = 119 Let, z¥2 denote the principal value of the square root. Sab def2?? 2) along the upper half of the unit circle (lz| = 1. b) along the lower half of the unit circle (fz = 1). Using the Cauchy theorem, Cauchy integral formula, or their consequences, evaluate the following integrals, all taken around the circle [z| - 2: a) fa, b) fa, °) $. ry fut ee) batts 5 $f 2B ae, i+? sinz ote Compute the integral eon Evaluate the integral a) given that C is the circle Jz — 1] = 1, b) given that C is the circle [z ~ i] = 1, c) given that C is the ellipse «7/4 + y?/9 = 1, x + iy = ¢) given that € is the curve x# + y4 = 4. (Hint: Describe the curve € in polar co- ordinates and observe that the problem hinges on the investigation of maxima and minima of the function f(@) = cost 8 4 sin* 8).} Consider the multivalued function z+1, F(z) = 7-1 Construct a branch /:(2) of F(z) which is continuous everywhere except for a cut along the real axis from z = —1toz— +1. Specily the branch carefully, i.c., give exact formulas for calculation of the values of /1(z) for all values of z. Determine the values of linn MAK ~ 1) + AG + ie). Constyuct another branch f2(z) of F(z) given in the preceding problem which is con- tinuous everywhere except for two cuts along the teal axis, from — © to —1, and from +1 to -hoo. Again, specify fa(z) with uumost care. Show that among the following complex series of constants, 2) ye a » > etn, Laan A ort, oy ott ta, saa logn the series (a) and (b) converge absolutely, the series (c) converges, but not absolutely, and the series (d) diverges. 120 FUNCTIONS OF A COMPLEX VARIABLE 29. Show that the following complex power series are absolutely and uniformly con- vergent for |2{ < 1: od (5! . 0) 3 FS teosme + finn), 30, Show that the series 1 (2*/n!) is absolutely convergent for all z. Show that it is uniformly convergent for [z| < R for arbitrary R. Despite the fact that R is arbitrary, why would it be incorrect to say that the series converges uniformly for all 2? 31. Show that the function f(z) = arclanz must possess a Taylor series about the point z = !. Without evaluating the series, show that its radius of convergence must be equal to 1/2. 32. Expand the following functions in a Taylor or a Laurent series valid in the im- mediate neighborhood of the specified points. Determine the regions in which the expansions are valid. a) f@) = cosz (2 = 0). b) f@) = @=0). aa + 2@2 + 1 @=2). 4) f@) = ©) f= wks @=D. ©) f@) = W@ED {2 = 0). [Hint: Multiply the series for e* and 1/(2” + 1).] D) f@) = coshz (2 = ix). 33. Find aif Laurent (or Taylor) expansions of the following functions about the indi- cated points: 1 VIO - way @ = 0), @-), b) fz) = 1 @+iF ) f= zig &=D, a) f@) = zt+1 e) f@) = log +z) & = 0%), f) fle) = Log = 0). Since the-last three examples involve branch lines and (possibly) multivatuedness, state cledrly which branch you use in each particular expansion. 34. Determine points where the following functions are not analytic. If the singularities are isolated, determine their character (if a pole, state the order) and calculate the residues: +1 Pad 1 YVO-F 5 WIO-EPPETH’ Yr : d) f= a e) s@) = tanhz, ) f@ = COs z conve 7 3 1\.. 3 3 1\. 3 w sey =(3~!)sinz — cos, wy ser= (3 -!)eose - Raine PROBLEMS 121 35. Show that f() = sin has an essential singularity at z = 1. 36. Using the residue theorem, evaluate the following integrals around the indicated contours: a) EEA $ 7 where C is the circle |z — 2i} = 1, cosh xz A : b) is Fry — Where C is the circle fz] = c) 4 ze!" ds, where C is the circle |2| c z~1 . sy ae d) cab $a" iz, where Cis the curve x + y° = 4. [See Problem 25(d) above.] 37, Illustrate the use of the residue theorem in evaluation of the following real integrals: 2 ao ax of, Treat” Fae (<9 of" 5 yi Fy or, T= feos ~ lo TF sink 5 2 dé 2xa a) [ @ Phan ni @— my > [OD os of cos" 6 dd = 123-5” Dan = 2" (n = even intezer). 0 2:4-6---n © PAG ZF Establish the following results on the basis of the techniques of Section 2.12: 40 3 2 x 7 dx a wf aan ° J, EIR 8 of” wota- fe tooska > 0; 5>0), ) f xsin xd Df wer D [watts cos x de GF aG? Fo) ~ {What is the result of (f) if'¢ = 57) Quote all theorems needed to justify the validity of your procedure. $22 FUNCTIONS OF A COMPLEX VARIABLE 39. Prove the following statements: 4° cosh ax a ay £ ASE ae = sec (lal <2) [Hint: Reduce to Example 6 in the text.] te wf. Tre" Ger (lal < 1). [UHint: Use the rectangle with sides y — 0, p = 2x, x = —R, and x = +R for the contour.] ° a - an Of Apt ie Z O Coch Oy + cos Bx can serve as a real part of some analytic function f(z). Evaluate o(x, »} = 1m f(z) by a method of your choice. Express f(z) explicitly in terms of 2. Al. The complex transformation w = be'*, where a, 6 = complex const, maps the isosceles triangle with the vertices at z1 = 5,22 = 6+ 1,andz3 = b+ 1 + iinto some curvilinear triangle 7” in the w-plane. Illustrate the invariance of angles property of this mapping by explicitly calculating the interior angles of 7’ to be 90°, 45°, and 45°, 42. Show that the image of the circle (x — 1)? + y? = 1 under the mapping w = 2” is a cardioid p = 2(1 — cos¢), where w = pe’*. Sketch the cardioid and locate the point where the invariance of angles breaks down. Explain. [Hint: Express the equation of the circle in polar coordinates in the z-ptane.] 43, Show that the integral “ [% civrean does not exist. Show that the integral 11 +H te ! [rs : az does exist and is equal (o zero. [iat Show that the primitive function of I/x is log |x|, which is valid for both positive and negative x.1 : CHAPTER 3 LINEAR DIFFERENTIAL EQUATIONS OF SECOND ORDER 3.1 GENERAL INTRODUCTION. THE WRONSKIAN A great variety of physical problems is reducible to linear differential equations, in particular, to those of second order, These also occur in the treatment of partial differential equations, The general form of such an equation is 409 £2 + a) + clay = Don. All variables are, in general, complex but special emphasis is put on real equations. it is customary to divide the equation by A(x) [points where A(x) vanishes must be given special treatment]: BY + PQdy' + Ody = ROD. If R(x) = 0, the equation is homogeneous. Otherwise it is nonhomogeneous. During the discussion of basic properties of the solutions, it will be tacitly assumed that the functions P, Q, and R are aualytic within the region (of variable x) undet consideration. However, P, Q, or R may {ail to be analytic at some particular points, e.g, when A(x) = 0. Whenever this occurs, special attention will be given to this fact. Let us first consider the homogeneous equation bY + Plxy’ + Ody = 0. A fundamental property of its solutions stems from the linearity of the equation: A linear combination of two solutions is also a Solution, namely, P= CQ) + Cayelx) is a solution provided y, and y2 are solutions. Function p(x) = 0 is always 2 solution and is known as the irivial solution. Two solutions are called /inearly independent provided that neither of them is a multiple of the other, Le., that the condition CiyiQ) + Coya(x) = 0 can be satisfied for all x only if Cy = C, = 0, Itis obvious that neither ¥. nor Je can be trivial if they are linearly independent. 123 124 LINEAR DIFFERENTIAL EQUATIONS OF SECOND ORDER 3 Linear dependence or independence of solutions can be established from the knowledge of their values and the values of their first derivatives at some particular point x = xo. IE yiG%o) = Kyalxo) and —-yi%o) = Kya(xo) (k= const), then y; and ys are linearly dependent. One way to see this is to note that the DE* yields ViC%o) = ky (0)- Now differentiate the DE and deduce thatt p7’(x0) = Ky"(xq), and so on. It follows that the functions y, and yz must be multiples of each other. Conversely, if ilxo)/¥olxo) ¥ Vilto)/ yao), then the functions must be linearly independent. To demonstrate this, let us introduce the function WiC), YAM = YrC¥2) — ¥e@WICD which is called the Wronskian of the two solutions y, and yg. If the two solutions are linearly dependent, then their Wronskian vanishes identically. The Wronskian has a remarkable property: Either it is identically zero or it is never zero; it cannot vanish for any particular value of x. To show this, calculate the derivative dw yt “ a ae 7 ve t+ vive — vay — vay = Yue — Jaye From the fact that y1 and ys are solutions of the DE, namely, Vi + Pdi + OCD” = 0, ye + POOvS + OG)y2 = 0, it Follows that Yiye — part + POW ys — yayil = 0 if we multiply the first equation by —y2, the second one by yx and add. Therefore dw a t Pew =9 and W(x) = Wlxo) exp {— {* PC) deh > (a) = Wire) exp {— J" C8) de} i; where W(x) is a constant, equal to the value of W at some particular point x9. * We shall use the abbreviation DE for differential equation. : 4 We assume that y(x) is analytic at x = x1, see pp. 129 and 146. 3.2 GENERAL SOLUTION OF THE HOMOCENEQUS EQUATION 125 Since the exponential function never vanishes, W(x) is cither identically zero (if Wo) — 0) or it is never zero (if W(xo) # 0), as slated. Returning to the condition yi(xo)/yo(x0) # ¥(x0)/¥o(x0), we sce that, in this case W(x) > 0 so that the Wronskian can never vanish, much less be identi- cally zero. Then y, and py, cannot be linearly dependent. It follows that they are linearly independent. Remark. The Wronskian is often represented in the form of a determinant; vee) y2Q) VCD 20) Wh), ya] = det 3.2. GENERAL SOLUTION OF THE HOMOGENEOUS EQUATION From the previous discussion it is evident that a solution of our DE is completely specified if its value and the value of its derivative are specified at a certain point X = Xp. From this, in turn, it follows that the formuta IO) = Cry) + Cry, where y, and yo are two linearly independent solutions and C, and C, are suitably chosen constants, represents the general solution of the homogeneous equation. Indeed, let p(x) be a solution with specified values Px) = 4, "(x0} = Substitute these values into the above formula and into the formula for y/(x) (obtained by differentiation): Crys(xo) + Caye(vo), b = Civilxo) + Covelxo). This system can be solved for C, and Cz and it definitely possesses a unique solution because the determinant det JilXo) YalXo) = Wo) Valeo) ¥2(%0) cannot be zero (y, and yz are Jincarly independent). If one sotution »4(x) of the DE is known, then a second one, linearly inde- pendent of y1, can be found by the following straightforward procedure: Observe that (2) 2s = Yeyt Wx) a\y, ye Dior 126 LINEAR DIFFERENTIAL EQUATIONS OF SECOND ORDER 3.3 The Wronskian is a known function: + = Wxo)exp j— | P(E) di W(x) = WOoderp {~ f" P(e) dt} except for the constant W(xq) (x9 may be chosen arbitrarily). Consequently, the ratio y2/y, can be evaluated up io a constant factor yal) _ v1) = = Woof op [- f Pode ists where both xq and x, are arbitrary. This implics that yo(x) is determined: Since any (linearly independent of y;) solution will suffice, the constant W(x9) may be chosen arbitrarily. The above result can also be obtained by the so-called method of variation of constanis,* due to Lagrange: If p(x) is a solution, then Cy (x) (C = const) is also 4 solution, but linearly dependent of y,; now try to find a Jinearly independent solution by making C a variable, j.e., assume that Pol) = COC). Substituting this formula into the DE obtain a DE for C(x), which can be solved in a straightforward fashion. The result reads ca) = af a [fread] th. where all three constants x, x1, and A may be chosen arbitrarily. Exercise. Give all the details in the derivation of the above formula for C(x). 3.3 THE NONHOMOGENEOUS EQUATION. VARIATION OF CONSTANTS ‘The linearity of the nonhomogencous equation allows us to state the following important theorem: If w, is a solution of the nonhomogeneous equalion and y is a solution of the corresponding homogeneous equation (obtained by setting R(x) = 0}, then the function u(x} = wrx) + vy) is also @ solution of the nonhomogeneous equation. The proof is trivial (by substitution into the DE). This theorem is the basis for the most common method of solving nonhomoge- neous equations known as the method of complementary function. Let u, ,be an arbitrary solution of the nonhomogeneous equation (known as the particular integral) and let p be the general solution of the homogeneous equation (known * Also widely known as the method of cariation of parameters. 33 ‘THE NONHOMOGENEOUS EQUATION 127 as the complementary function), Then u = uw, + y represents the general solution of the nonhomogeneous equation. Indeed, 2 must be a solution by the preceding theorem. To show that it is a general solution, let us represent the complementary function as a linear com- bination of two linearly independent solutions of the homogeneous equation. Then ux} = i) + CiyiGd + Coyolx). If we set u(x9) = a, and (x9) = 6, then the following equations must hold for Cy and Cy: Cryrlxa) + Coya(xe) = @ — inbxo), Cyilro) + Caph(xo) = 6 — wiry). Since W(y1, ye) > 0, this system has a unique solution for C, and C2, and the theorem follows. A particular integral (or the general solution, for that matter) of a nonhomoge- neous equation can be obtained in a straightforward fashion if the complementary function is known. The particular integral is sought in the form wa) = or)y(x) + vay o(a), which can be obtained from the complementary function by replacing the con- stants C, and C, by the functions uv, and vg; hence we have the name variation of constants. Note that the above relation does not define (x) and v(x) uniquely, even if u(x) is specified. For instance, it is possible to add to vy(x) an arbitrary function a(x) and at the same time subtract from v(x) the function BOOP 1CO/y 200. because (01 + aiys + [on — gys/valye = viyr + vayr. We can profit from this arbitrariness by imposing certain additional requirements on vy and v2 to facilitate their determination. Let us calculate w’(x): , w= vy + vaya + viv, + vby2- If we make the additional requirement that viv + voy. = 0, we simplify the determination of o, and v2, as seen below. This also has an inter- esting practical meaning: If the value (xg) and the slope w/(x9) are fixed at some point xq and if u(x) were to be approximated near xq by a linear combination Ciyr + Cops, then the constants C, and Cz would be determined from the equations uo) = Ciys(%o) + Copa(xo), u'(Xo) = Ciyi(xo) + Coyalra). 128 LINEAR DIFFERENTIAL EQUATIONS OF SECOND ORDER 34 ‘These values of C, and Cz are known as osculating parameters of the solution u(x) at the point xo. The condition imposed on v4 and v4 implies that ux) = vi@)yr(x) + o2e)y2(2), W(X) = v1) + vay). In other words, »,(x) and v(x) are just the (variable) osculating parameters of the solution u(x). Now calculate w’’ and substitute u”’, w’, and % into the DE. Utilize the fact that y; and ye are solutions of the homogencous equation. The result reads iyi + vgys = RK. This relation, along with the requirement that Diy: + vay2 = 9, ields » oi = —yaR/W, oh = yi R/H, where W — Wronskian of y; and yz. Now obtain, by direct integration, nye —f BORO a, ogy = [ MORO a Here x: and xz can be arbitrary if an arbitrary particular solution (x) is sought. In fact, a different choice of x1 and x2 changes v; and v2 by additive constants so that the resulting solution u(x) is the general solution of the nonhomogeneous equation. : The results of the last three sections may be summarized in a conclusion: To find a general solution of the nonhomogeneous equation yi + Py + Op = R, it is necessary to find only one solution y; of the homogeneous equation py’ + Py’ + Qy = 0. Then the second solution y» (linearly independent of y1) of the homogeneous equa- tion can be found by direct integration and, after that, the general solution of the nonhomogeneous equation can also be found by straight integration. 3.4 POWER SERIES SOLUTIONS Many finear differential equations of second order have solutions which are not expressible in a simple form in terms of well-known functions, such as algebraic, trigonometric, and logarithmic functions. Some seemingly “simple” equations like y” + xy = 0 fall into this class, We may obtain, however, a valid repre- sentation of many such functions in terms of power series. 34 POWER SERIES SOLUTIONS 129 From the very nature of the second-order linear DE it follows that if P, 9, and R are analytic at some point x = a, then the solution y, if it exists at all, must be analytic at that point. Indeed, the DE implies that y(x) possesses first and second derivatives in some neighborhood of x = a {a neighborhood over which P, Q, and R are still analytic), Therefore, the solution can be expressed by a Taylor series: YO) = co + €x(x —~ a) + eax — a)? + 03(x — af 4 --- The most familiar series of this type is the Maclaurin series (or power series in x) which is, of course, a Taylor serics for the special choice a = 0, Many differential cquations will have a solution analytic at the origin which is then expressible as IO) = cob crx + cox? + egx2 4 ++ OF course, this is true only within a certain radius of convergence R. Bul, within this radius of convergence, the series can be differentiated any number of times, so that p’ and y” are expressed by similar serics, This forms the basis for the power series method, illustrated by the following example: Consider y” — k¥y = 0 (k = real const). (The general solution of this equation is Cye** + Cae—*, so that the result can be checked.) Assume the solution in the form YOR) = ey + eyx + eax? + cgx3 + --- Then a YO) = 4 + 2eax + 3cgx” + deyx? +--+, YC) = lea + 3+ Degx $ 4+ Bex? fo The series for y” and y are now substituted into the DE and the coefficients with all powers are set equal to zero [on the basis of the theorem that all Taylor co- efficients of the function f(z) = 0 must vanish]. It follows that Pea ~ k?eg = 0, 3+2ey — key = 0, 4+3cq — K%eq = 0, and soon. This yields Pa B 2 , eg y fo eg = zy c= 4.3m etc. In general, it is seen that Pe ae Such a formula is called a recursion formula (or recurrence formula) for the co- efficients. If ¢9 is Known, then all even coefficients can be determined. Odd coefficients can be determined if ¢; is known. The entire solution splits into even and odd parts and can be writlen as HOD = COL PRE ERR RD cae FREY RB/SE ) cn (allt > 2). 130 LINEAR DIFFERENTIAL EQUATIONS OF SECQND ORDER 3.5 Both series appearing above converge absolutely for all values of x (cg., by ratio test) and uniformly within an arbitrary radius of convergence (by Weier- strass M-test). Consequently each one of them represents an analytic function over the entire complex x-plane, and both functions evidenily satisfy the DE. Therefore, we have obtained the solution in the form HX) = Coy) + crv2), where (x) and po(x) are solutions, and co and c, arc arbitrary constants. ‘The solutions y; and ya are obviously linearly independent since one of them is even and the other is odd, We conclude that we have oblained the general solution of the DE. Jn our particular example the series are recognized as hyperbolic functions: ya) = 3 + 2x /21 AAA Vol) = x + 2x3! kAxP/5! ‘The general solution p(x) = co cosh kx + (€; /k) sinh kx is, of course, equivalent to p(x) = Cre + Coe. Remark. The solution (x), being an even series, implies that »',(0) = 0, while the solu- tion y2(x), being an odd series, implies that y2(0) = 0. cosh kx, = (1/k) sinh kx. 3.5 THE FROBENIUS METHOD The power series Eno cnx" is a Taylor series about the point x = 0. It will represent a solution provided the solution is analytic at x — 0. If the solution is not analytic at x = 0, we may, of course, switch to some other point x= a where it is analytic. However, there is another approach: If the solution is not analytic at x = 0 because it has a pole there, then it will possess a Laurent series of a similar form: y@) = y nx” except that it will not begin with» = O but rather with 2 = —-#1, where m = order of the pole. The method of Frobenius utilizes this idea: We seek a solution of the DE in the form ' = x cnxt” (Co 9), a0 yx) = x* z where s is left completely undetermined. This generalized power series is capable of describing , a) analytic functions that do not vanish at the origin (s = 0), ' b) analytic functions with a zero of order m at the or = m = positive integer), ~ ¢) functions with a pole of order m at the origin (s = —m = negative integer), 4) functions with certain types of branch points at the origin (5 = noninteger). 35 THE FROBENIUS METHOD 131 An example of (d) is the function f(x) = \/x/sin x which is expressible by the series f%) It XU + X16 + phox" + rftenx® +0 = PU A bet + abet to). If the solution of a DE is expressible by the series {co ¥ 0), then, in general, such a series can be differentiated term by term. Indeed: L. If s is zero or a positive integer, the series is a Taylor series and can be dif- ferentiated any number of times within its radius of convergence. 2. If sis a negative integer, the series is a Laurent scrics and can be differentiated any number of times within the annulus for which it holds. 3. If s is not an integer, then p’(x) is expressible as yO) = @Y Dax $0 (= cnx"): y n= The expression x* means, of course, some particular branch of x’ (recalt the definition x* = e’}"*=), It is not difficult to show that, for any branch, = sxot except at the branch line for this particular branch, More precisely, do atone) _ fab tose ae =e with the same branch of Log x on both sides of the equation. In short, x* can be differentiated by the same rule as x* {1 = positive integer). This implies that the whole Frobenius series y@) = z ext? a0 Figure 3.1 can be differentiated term by term. At worst, the results will be valid in a region fike the one shown in Fig. 3.1. The same statement is true for »”(x) and y’(x). The series for y’(x) and y”(x) are now substituted into the DE and the co- efficients of equal powers of x are equated to zero by similar considerations as for power series. The resulting relationship often permits the evaluation of one or two solutions of the DE. Applications of this method and the new problems raised by it are illustrated by the following examples. 132 LINEAR DIFFERENTIAL EQUATIONS OF SECOND ORDER 35 Example 1. Pole at the origin. The DE reads xy! day! + (x? — Dy A solution is sought in the form v3) =D ea & Then YO) = x en€s + matte, VOD =X cals + aye ean — I. nao Substitute into the DE, and obtain BS leas + as +o — WF ens +) — eb + DL gx? = 0. aad n=O To compare the coefficients with various powers of x it is convenient to standardize the notation so that the powers of x have the form x°*" in the second sum, as well as in the first one. For this purpose, set 2 + 2 = n’ in the second sum and obtain xd eggs nme Since n’ is a “dummy index,” it may be “unprimed” and this sum may just as well be written as ¥ cao a The DE then assumes the form E alls + myst — aT + Z cae = 0. x nat Write separately the powers of x which appear in the first but not in the second sum, namely, the terms with n = Oand a = 1: eds(s -+ 1) — 28 + calls + 1s + 2) — Itt +E fells + ls tb D4 Gad 5 0. Paar Ttis necessary, first of all, that eols(s + 1) — 2) = 0. 35 THE FROBENIUS METHOD 133 Since cy > 0 (x* is assumed to be the lowest power of x appearing in the series), it follows that ss + 1) 2=0, Such an equation, arising from the lowest power of x on the left-hand side of the DE, is called the indicia? equation; it determines possible values of s. In this particular case these values are s, — 1, 52 — —2. The second condition to be satisfied is alls + DG+ 2-7 =0. Whether s = t or s = —2, it follows from this equation that cy must be zero, The next condition, valid for all values of 7 greater than or equal to 2, reads calls + als + t+ 1) 21+ G2 = 0 or 1 =CEmetaE He? 22. This recursion formula determines all other coefficients (note that the denominator cannot vanish if a > 2). For s = | we have ! 1 2 = FZ 4 = — 79% ca i Eo = she etc, 4 75g? ~ 30°" 9 and, evidently, eg = cg = ++" = 0, This yields the series PIC) = en(x — x3/10 + 5/280 — ++) This series is absolutely convergent (by ratio test) for all values of x in the complex x-plane and defines, therefore, a function which is analytic for all (finite) values of x. This function has a simple zero at the origin. Now take s = —2. Then Again, ¢y = ch = +++ = 0. The series is vox) = c0(I/x? + ~ 7/8 +--+). This series is absolutely convergent (by ratio test) for |x| > 0. It is a Laurent series and defines a function that is analytic for all x except x = O where pox) ol 134 LINEAR DIFFERENTIAL EQUATIONS OF SECOND ORDER 35 has a pole of second order. In this case the method of Frobenius has led us directly to the general solution of the DE. Remark. If we setco = 4, then the function y:(x) actually becomes dr@) = sin x/x? — cos x/x, and is known as the spherical Bessel function of order one, Similarly, if = —1, then ya{x) becomes mix) = 00s x/x? ~ sinx/x and is known as the spherical Neumann function of order one. Both of them, along with other related functions, are of great importance in physics. Example 2. Branch point at the origin, The DE reads xy" + xy’ + (x? — Dy = 0. Write, as before, ya 2D eax, tno vn z ents + nxt te-§, Eo ys x cals + ays +n ~ 1)Pt*-?, ano and obtain DS calls + aye tom — 1+ (tm) Be tL cox? = 0, nad n=0 The indicial equation (coefficient with x’) is s(s — 1) + 5 — 4 = 0, yielding the roots sy = 4,52 = —}. Consider s = 4; then e,{3- 4+ 3 — 4] = 0, so that cy = Use a -+ 2 = nin the sum Lae Cxx?t"*?, combine this with the first sum, and deduce the recursion relation el@ + PO — D+ + DHF tae = 9, or om = (@> 2). 1 ED It follows that all odd coefficients vanish and the series reads 1 2 x i JQ) = Cox 1-354 - ‘ This scrics converges for all x (by ratio test) and yields a solution that is analytic in the whole plane except for the branch cut of 1/2. 3.5 THE FROBENIUS METHOD 135 Now consider the second case, s — —}. The equation for ci now reads el(-4) + 4 ~ 41 = 0 and is satisfied by all values of c. In other words, c, remains arbitrary. The recursion formula is 1 me Ho aan (a> 2). Since A need not be zero, it is no longer true that all odd coefficients vanish. The in general, contain both even and odd powers of x and can be written as va) = cx FUL — 22/2 + x49/Q-3-4)— etx x — 8/23) + x82 3-4-5) — This series also converges for alf x. The branch line for x"? can be chosen to be the same as in p(x) so that both solutions are analytic in the same region. It now appears that y2(x) alone represents the gencral solution since it has two arbitrary constants multiplying two linearly independent parts. (Note that one series is even and the other is odd, so they must be linearly independent.) Therefore, the solution y,{x) must be a special case of yo(x), and indeed it is: Set c4 = 0, c} = ¢o in ya(x} and obtain y4(x). Remark, It is not difficult to sec that this phenomenon (one root of the indicial equation determining the general solution) can happen only if the roots of the indicial equation differ by an integer. It is then the smaller root which yields the general solution. The solution y2(x) obtained above can be written in a more compact form as 12 yolx) = ex"? cosx + efx! sin x. The functions Jajalx) = V2 x7"? cos x and Fue) = VT ]x xo? sin x are known as the Bessel functions (of the first kind) of order — }and 4, respectively. The solution can be written as Yalx) = AS sa(x) + BI 112). Remark. The DE x2y’ + x + (x? — vy = 0 ( = nonnegative real number) is known as the Bessel differential equation of order y. Us solutions, known as cylindrical functions (or Besset functions of various kinds), occur in many physical problems, Example 3. Convergence questions. Consider the DE (1 = x?)y'? — 2xy + ay = 0 (A = real constant). 136 LINEAR DIFFERENTIAL EQUATIONS OF SECOND ORDER 35 The Frobenius series = y= Do eg" a=0 leads to SS cals + ahs ton — Ux? a= - Z olle + mis +n — 1) — s+ a) + A = 0. The indicial equation stems from the lowest power x¢~? and reads s(s — 1) = 0 yielding the roots 5; = 0, s2 = 1. The equation for ¢, (power x*~") reads c,(s + Us = 0 and yields cy ifs = 1, and arbitrary c; ifs = 0: the case discussed in Example 2. The recut formula is -@-Da-) ape ED The root s = 0 gives rise, therefore, to the general solution in the form Wx) = co Deke" er Dax" neweren anodd The root s = | yields, evidently, the odd part of this solution. Now consider the convergence of the series. By ratio test, (2 — 2X0 — 1) — n(n = Ty) |_eax"_ ha n-| Both series converge for |x| < 1, whatever the value of & may be. For |x| = !, the ratio test is inconclusive. However, the integral test yields ME at Me @-M-D-a,_ ["@-D x [ at — 1y a -/ 7 -f ac nt and since a f 24> 0 as M— oo, both series, in general, diverge* for [z| = f. There is, however, an exception: If A happens to be of the form \ = K/ + 1), where / is a nonnegative integer, then * This is not unexpected: Our equation, when written in the form y” + Py’ + Oy = 0, reveals that the coefficients P(x} and Q(x) ere not analytic at x = 1, It is not un- reasonable to suspect that the solutions may also fail to be analytic at these points. In this case our series (both are Taylor series) will indeed have a radius of convergence equal to unity. 35 THE FROBENIUS METHOD 7 one of the series will terminate giving rise to a polynomial. The other series will be divergent. A similar statement holds for |x| > 1. In conclusion, the Frobenius method yields a general solution of the DE only for the region |x| < 1. However, if\ = X/ + 1), then there is a particular sokition of Frobenius type (namely, 2 polynomial) which is valid for ali vatues of x. The DE considered above is known as the Legendre DE and is one of the most important ones in physics. In particular, physical conditions usually require that the solution be analytic at |x| = 1. Such sotutions arc possible if A = I(F+ 1) and are obtained from the general solution V(x) = Co . cnx” ey S, cnx” by setting c, = O(if lis even), or co = 0 (if lis odd), Itis, furthermore, customary to standardize the solutions by the following choice of the coefficients with the lowest powers of x: each = (—1)"7#— = "V = even), AT TT These standardized solutions are denoted by P;(x) and are called Legendre poly- nomials (or Legendre functions of the first kind). The lowest ones are Pol) = L PiQ} = x, Po(x) = 46x? — D, Pa(x) = 5x" — 3x), Pa(x) = §635x4 — 30x? + 3), ete. @ = odd). ‘The second solution of the Legendre equation, linearly independent of Pi(x) and appropriately standardized, is denoted by Q,(x) and is called the Legendre function of the second kind. Functions Q:(x) (f= 9, 1,2,...) are multivalued and have branch points at x = -L1. For [a < 1, we customarily sclect the following definition for the lowest function: Ql) = TER HE RV ARTE (Id < or, in a compact form, Qo(x) = artanh x = flog re dal <1). 138 LINEAR DIFFERENTIAL EQUATIONS OF SECOND ORDER 35 The other Legendre functions of the second kind are then expressed in terms of Qo(x) as follows: O10) = PiX)Oo() — 1, Q2(x) = Palx) Qo) — bs Qa(x) = PalQo(x) — $x” + §, ete. The functions Q;(x) are also used in physical applications but to a far lesser extent than Legendre polynomials. Example 4, Logarithmic singularity. The DE reads xy’ + y' = 0. The Frobenius series p(x) = Sofuo cnxt*™ yields x leas + sta N+ alt np =9 m0 and cas +a)? =0 (ln). The indicial equation is cgs? = 0 and has the double root s = 0. It is seen that there is no recursion formula (in the proper sense of the word). Rather, co may be chosen arbitrarily, but all other coefficients vanish. ‘This results in only one solution of Frobenius type, namely, a(x) = co = const. It is reasonable to conclude that the second solution cannot be represented by a series of Frobenius type. Possible causes of such a failure may be that the second solution y2(x) possesses at the origin (a) an essential singularity, or (b) a branch point of a special type, not covered by Frobenius series. Whatever the nature of the difficulty at the origin, the second solution must be analytic somewhere and a Taylor expansion about another point may be at- tempted. An alternative procedure is to utilize the Lagrange method (of variation of constants) mentioned in Section 3.2. In this particular case we have P(x) = t/x, so that (set yy(x) = ¢p and choose x9 = x; = 1, A = eo) yale) = Clxleo = deo f exe(- [2 a 4 _A fd _A _ Af. op LOB * = Log x. This solution is, of course, a multivalued function, but it is sufficient te select some (arbitrary) branch, say the principal branch ys = log x, to serve as our’second solution: All other branches of Log x can be represented as linear combinations of the two basic solutions y,(x) = I (we have chosen cy = 1) and yo{x) = log x. . | 35 THE FROBENIUS METHOD 139 Our conclusion is that branch points of logarithmic type may be encountered when solving some simple linear DE of second order. This may suggest that we try, in general, series of the type Jog x Y) cax™™ — (eg = 8), amd or perhaps, the combination yx) = log x Do eax + ax (co 0, ag ¥ 0). nao mao This type of expression may be called a generalized Frobenius series and may be used to extend the Frobenius method. It is very common in physical applications. Example 5. Double root of the indicial equation, Consider the DE ay" + y+ xy = 0 which is known as the Bessel DE of order zero. The Frobenius series v(x) = DS cart" — (en ¥ 0) Fo results in the indicial equation, cos? = 0, the equation for c,, c(s + 1)? = 9. and the recursion formula, a= @2 9). ! Cryo? Analysis shows that there is only one solution of the Frobenius type, i.e., the series of even powers of x: n= LD CH? x" (with OF = 1). Tn neOTA ie (8) G!) Exercise. Spell out the detailed arguments leading to this conclusion. The series converges for all values of x. If the standard choice co = I is made, then the solution is known as the Bessel function of order zero: hio= > CW in nmoven (0 a(S An attempt will now be made to find a second solution in the generalized Frobenius form: v(x) = log x x, eat tt ot > gx (co # 0, ay * 0). r= 140 LINEAR DIFFERENTIAL EQUATIONS OF SECOND ORDER 35 Before p(x) is differentiated and substituted into the differential equation, observe that the series X ex" =o cannot possibly be anything but Jo(x) (or a multiple of it). Indeed, write YO) = log x- u(x) + 209, where u(x) and o(x) are Frobenius series. Then y= wlogxt (i/xa to, yf? = ul log x + (2/x)u! — A /x a boot Substitute into the DE: q log x(xu” 4 a! + xu) + Qu! + 20" 4 vo! + xv) = 0, Ail terms in parentheses are Frobenius series and since log x is not expressible as a Frobenius series, it is not difficult to conclude that xu" + ul + xu = 0, which proves the statement since Jo(x) (or a multiple of it) is the only Frobenius- type solution of the DE under consideration. This simplifies the problem: It is now necessary to find 2) = Eo aax"™ (an #0) satisfying x0” + v’ + xe = ~2j. Write Folx) = x b,x", where the coefficients 5, are known. Then 80) = FS ume Substitute this expression, along with the Frobenius series for », v’, and o!’, into the above equation. Obtain, in the usual fashion, / agree a PH + Lal FY! gal = FS 22nd Axo 35 ‘THE FRODENIUS METHOD 141 Recall that zg (n= even), (a = odd). ‘The lowest power of x on the right is, therefore, the first power. Consequenitly, the choice r = 2 may be tried. Then Gg = —(2-2- by)/2? = —by = f. Also, a, = 0, and Ont 4 2)? + ay 2 = — Am + Bmae CD 2. The last relation yields the recursion formula 1 2 On = ~ Gp apt ep gins (mE), from which all other coefficients a, may be calculated (in particular, all odd coefficients vanish). This yields ox) = DY anx™? = pen gett poe A 733 and establishes the second solution as Pa(a) = log x: Jo(a) + 0). ‘This function is denoted by many authors by Yo{x) and is called the Bessel function of the second kind of order zero. Remark. In modern literature it is customary to select a different function to represent the second solution, namely, the following tinear combination of Jo(x) and Yo(x): Nol) = (7/2)¥ 0x) 4+ (Y — tog Dol), where ¥ is the so-called Euler-Mascheroni consiant: Y= lim 35 —1oe n) - tim ( theta t4— tem) = 05772... Ware eT Pe Nowe 2°03 N The function No(x) is usually called the Neumann function of order zero. However, there are authors who denote this function by Yo(x) and call it the Bessel function of the second kind of order zero. In general, the nomenclature and standardization of special Functions like these are far from being settled. Example 6. Complex roots of the indicial equasion, Consider the Exler DE: yt xy ty = 0. 142 LINEAR DIFFERENTIAL EQUATIONS OF SECOND ORDER 45 The Frobenius series = yO) = DO cyt" n=O yields the relation cal(s + 1)? + 1] = 0. The indiciat equation (set 2 = 0) cols? + 1) = Ohas the complex roots s = +7. For either of these two values, all coefficients c,, (except ¢o) must vanish. If cy is chosen to be unity, two solutions are obtained: BiG) = x ya) = x8 Both functions are multivalued. Select the principal branches: vil) = eR, ig) = ethene, The Wronskian of these two solutions reads Wy, yo) = ebtonze Hon s( 5/5) — eitonagilon eee yy = ~ifx~ ix = ~2ifx. Since the Wronskian never vanishes (for finite x) the solutions are linearly inde- pendent and the general solution is obtained: 9) = Ae !8= 4 Bemilose, where A and B are complex constants. Ifa real solution is desired, write gilorr _ gilfonteltiarge) For real and positive x, arg x = 0, and log (x| log x so that etlonz = cos (log x) -+ isin (log x). The real functions cos (log x) and sin (log x) are linearly independent, and each must satisfy the DE (see the Exercise on p. 52). The general real solution may be taken as YX) = €1 c0s (log x) + casin (logx) (x > 0). This solution is meaningless for x <0. However, a different branch of x‘ may be selected fo produce the general solution in this case as well. Observe that both solutions have 2 branch point of logarithmic type although they are formally represented by a (terminating) Frobenius series. Example 7. Roots differing by an integer. Consider the Bessel differential equation of order one xy" $ xy + (x? — Dy = 0. ' The Frobenius series JX) = 3.5 THE FROBENIUS METHOD 143 yields the indicial equation, co(s? — 1) = 0, the equation for ¢1, e1s(s 4-2) ~ 9, and the recursion formula, I ~Ghatieta— ye 62? a = The roots are s = +1. In either case, cy = 0. However, an attempt to construct the Frobenius series with s = —1 fails because cz is not defined by the recursion formula. Colloquiatly speaking, cz “blows up” (or “becomes infinite”). The second root, s = -++1, transforms the recursion formula into = ones @> 2, "= — apa wae ee which readily determines all coefficients. The choice eo - $ yields the Bessel function of order one: Ji Sy 4 wtta dy det ghgata newton r(51) @ +2) ‘The second solution can be attempted in the form (as in Example 5) Ya(x) = log x- AG) + vo) and leads to the DE for o(x): xo" + xe + G2? — No = — 2x). Since J,(x) has a simple zero at the origin, it is convenient to write SiG) = 3% byt! =o Then ~ wai) = 3 (tae + DH n=O Assuming u(x) to be a Frobenius series, v(x) = 0 anx't" — (ap ¥ 9), sao and substituting into the DE, we obtain aot + WG — Wx" HE ayr(r + 2)x7* + Lane tb Ye ot — VD beatae maa = CHG t De ai 144 LINPAR DIFFERENTIAL EQUATIONS OF SECOND ORDER 3.5 Recall that (py? = even), = add). so that the series on the right-hand side has only odd powers and starts with the first power. A value for r should be chosen so that the series on the left-hand side has the same form. A little reflection indicates the likely choice: r = — 1, a, = 0. It would appear, at the first glance, that ao is arbitrary. This is not so, however. The equation now becomes De [em ~ 2) 4 amadde™! = D> [-2on(n DE! ay =o Equate the coefficients with the first power of x on both sides (m = 2,n = 0): 9 = —2by = — 24) = -1, Observe that a2 is not involved in this equation and remains arbitrary. To find other coefficients, set m = + 2 and obtain the general recursion formula G4 an(n + 2) + a, = —2b,(2 + 1). The case m = 0 has already been considered. For 2 > 0 this relation can be solved for a,42: | 1 Med) Ont = — ie ED) — nab 2) It follows that all odd coefficients vanish, since a, = 0 and all odd 6, vanish. Hf az is selected, then all other even coefficients can be calculated, It is not difficult to see that different choices of az amount to adding to v(x) a multiple of J(x). This is not surprising since if the function bn (a > 0). u(x) = fog x Ji(x) + ey) is a second solution (linearly independent of J;), then u(x) + iQ) is still a solution, linearly independent of Ji (x) and having the same generalized Frobenius form. One customary choice is a2 = —4, which leads to the function 1 I VQ) = fog x) — 2 gx t 5px" - 5 = ; {known as the Bessct function of the second kind of order one). Another choice of the second solution is the Neumann function of order one: N(x) ~ Q/a)EY Gx) + Or — log 2Q)} 35 THE FROBENIUS METHOD 145 Example 8. Three-term recursion formula, Consider the DE py” + (1 — x7)p = 0. If yO) = Dee cnx? *", then DY G+ ays toa — Vea? + eat — YE cgxt et? = 0. nao “=o no The indicial equation cqs(s — 1) = O has two roots, s = O ands = 1. Consider the root s = 0. The equation for ¢, reads cx( + Is — O and leaves cj arbitrary. For the next two coefficients (1 = 2 and = 3), the second sum is involved leading to n(n — Wen + Coa = 0 or nna ey = 2,3). Forn > 4 the third sum becomes involved, and we have n(n — Wen + Cr-2 ~ Goa = 0 Obsetve that the general recursion formula involves three terms. It is scen that if ¢ is chosen, then cg = —co/(2‘ 1) = ~4ep and cq is now calculated from co and cy. Alf other even coefficients are obtained in a simitar fashion and are independent of the selection of odd coefficients. It is possible then to set c; = 0 and to construct an even series solution since from c, = 0 it follows that cz = 0, and then cs = ¢7 = C9 = ++° = 0, The series reads (with co = 1) DQ) = bn bet pet a ae to and is easily recognized as mn (AG) HG) eo ‘The second solution may be obtained as the odd series, generated by the coefficient cy. We set ¢o = Oandc, = | (this is somewhat unorthodox since ¢o is, as a rule, assumed not to vanish; however, the ultimate results are not affected by this departure from the usual technique). Then the following series emerges: ya) = x at rhgt ada Fo Each of the series, y,(x) and y2(x), is a solution of the DE, and they are linearly independent (one is odd, the other is even). The general solution PO) = eur) + erya(x) 146 LINEAR DIFFERENTIAL EQUATIONS OF SECOND ORDER 35 is, therefore, obtained. As for the root, s = i, it is easy to verify that it will generate, basically, the solution y2(x). Remark. This DE occurs in quantum mechanics where it is used to represent the ground. state of a harmonic oscillator. Physical principles demand selection of the solution which approaches zero as x— Loo. The solution y1(x) satisfies this condition, whereas the solution y2(x) does not. This may not be evident from the series for y2(x), but by the method of variation of constants it is not difficull to show that yale) = yal) f ope” f ef at. The integral Jo et” d& represents the area between the curve y = e" and the x-axis. Bear- ing this in mind, it is almost trivial to show that, for instance, f eat>Xe" (> 0), D 2 which implies that [y2G0l > klen > asx] 0. The examples presented above show that the Frobenius method, when suc- cessful, leads to the solution of a second-order linear DE provided this solution is either (a) regular at the origin, or (b} possesses a pole or branch point of power or logarithmic type at the origin. Obviously, it is desirable to determine in advance whether a given DE will have solutions with these properties. Two basic theorems to that effect are given below. They involve the following definitions. Definition 1, if the functions P(x) and Q(x) in the DE y+ Py’ + Qo = 0 are analytic at the origin, then the origin is called ordinary point of the DE. Otherwise, the origin is called a singular point. Definition 2. 1f P(x) and Q(x) are not both analytic, but are of the form oo) = ¥&x) . Pa) = OW) = where g(x} and ¥(x) are analytic at x = 0, then the origin is called the regular singular point. Otherwise, it is called an irregular singular point. Theorem 1. If the origin is an ordinary point, then any solution of the DE is analytic at x = 0. ‘Theorem 2. Theorem of Fuchs. If the origin is a regular singular point, then any solution of the DE is cither (a) analytic at the origin, or (b) possesses a pele or branch point of power or “logarithmic’ type. (In othe? words, it is expressible as a generalized Frobenius series.) 3.6 SOME OTHER METHODS OF SOLUTION 147 Proofs of these theorems can be found in advanced texts on differential equa- tions or complex variables. Remarks 1, If the origin is an irregular singular point, there is still a possibility of a solution of the Frobenius type. Hf one such solution is found, the other may be obtained, for instance, by the method of variation of constants. 2. The entire theory outlined above can be applied to any other finite point by a trivial argument. Finally, here is a brief summary of possible situations encountered in the use of the Frobenius method: If the initial attempt to find a solution is made in terms of a Frobenius series, (co 0), then the lowest power of x gives rise to the indicial equation which is a quadratic equation in s (with real coefficients for a real DE). The following cases arise, according to the nature of the roots (which may be complex, even for a real DE). CASE 1. Two distinct roots not differing by an integer, There ate two linearly independent solutions of power typc. Each solution contains one arbitrary parameter (Example 6). CASE 2. A double root. There is one power-type solution and one logarithmic- type solution (Examples 4 and 5). CASE 3. Two distinct roots differing by an integer (two subcases are possible): a) Two solutions of power type, each corresponding to one root. The lower root yields a two-parameter solution. The higher root yields a one-parameter solution and is a special case of the first solution (Examples I, 2, 3, and 8). b) One solution of power type, given by the higher root. The lower root fails to give a solution (infinite cocfiicients), but a logarithmic-type solution exists (Example 7). 3.6 SOME OTHER METHODS OF SOLUTION Power series of ascending powers of x are convenient in the vicinity of the origin. For “far-away” points it is much better to use power series of descending powers of x (as in the principal part of a Laurent series). For instance, if a function f(x) is analytic at infinity, then it must be representable by a scries {= x m(1/x"} valid for |x| > &, where R is some nonnegative constant. Such expansions are usually called expansions ghout the point at infinity. 148 LINEAR DIFFERENTIAL EQUATIONS OF SECOND ORDER. 3.6 Expansions of this type may be attempted dircetly in the process of a solution of the linear DE of second order. However, it is advisable to proceed in an indirect way by making the substitution x = 1/z. Then dy _ dy de ix de des H(z ), ay _ dy edz ad’y vay, ae? -21(0) 9 feo at and the DE becomes fr -ar()lg+t0()y-* ‘The desired expansion in descending powers of x is now a Taylor expansion about z= 0. If the coefficients Fe=2-dr(!): «= £0(2) are analytic at z — 0, then such an expansion exists and can be found by standard methods. Example 1. The DE reads x*y” + 2x3y/ — y = 0. On division by x4, we obtain were by tyn0 The expansion y(x) = Drag OmX is desired. After the transformation x= 1/2, we obtain aate nt. I FQ)=2-hx=0, a@=-h Both F(z) and G(2) are analytic at z = 0, and the new DE reads ‘The desired series arc yi@) = yi) = aap at x Wx Wx n@= SE ewetar-tey tthe. 36 SOME OTHER METHODS OF SOLUTION = 149 Both series converge for |x] > 0 and represent the expansions about the point at infinity of the functions yi(x) = e%*, yo(x) = 7", Example 2. Consider the zero-order Legendre differential equation (1 — xp" — wy = 0. It is desired to obtain the expansion of Qo(x) about the point at infinity. Divide by (1 — x®)to obtain 2x tT 2? Here PG) = 2x/(x? — 1), OG) = 0, and ra) =2— zP()) = x jo GE) =O z, z 2 The new DE reads and has exactly the same form as the original one. It is known to possess the solutions 2? FGte Gl < Dd a r@=1 meat 5+ Therefore the original zero-order Legendre equation possesses the solutions 1 1 1 n©=t nosbtggtet ate dl. x The first solution is P(x). The second solution map be called Qo(x); it can be written in compact form as Qo(x) = arcoth x = glog (e+ DAK - DY) (xl > D. Remark. The function Qo(x) is multivalued. The choice of branches 1tx Q0lx) = Msraas He 1 Flog Lt, Ix] >I is, of course, arbitrary, but it is the standard choice. If a problem warrants it, another selection of branches can be made. ‘The above discussion deals with expansions in integral powers of x. It may evidently be extended to generalized Frobenius expansions about the point at infinity. 150 LINEAR DIFFERENTIAL EQUATIONS OF SECOND ORDER 3.6 Lineac differential equations of second order are often transformed, for various reasons, by a change of variable. The dependent variable y is usually changed by POY = rir), where y:(x) is known or determined by certain conditions; the result is a DE for o(x). An example of this technique has been found in the method of variation of constants: here y,(x) is one of the solutions of the DE. Other applications include (a) the elimination of the first-derivative term, and (b) the factorization of the behavior at infinity, If the above substitution is made, then the DE becomes yw + yi + Pye! + OF + Pri t+ Qyi)o = 0. Instead of eliminating the third term by choosing y; to be a particular integeal, we may require that ay + Pr = and eliminate the first-derivative term. This means that vil) = exp {—4f" PCE) ae} and the DE for » reads o — (2 — HP — APA = 0. This technique finds its applications in estimating the behavior of the solutions “at infinity” and various approximation methods. Example 3, Consider the DE xy” + 2xy’ + (x? — 2jy = 0 (spherical Bessel equation of order two). Divide by x7: + (2/x)p" + (1 — 2/x2y = 0. Let y = p10, where 1 f'2 1 a= oo(-3 / 2arl = exp {—log x} = ‘Then v satisfies bY + (= 2/22) = Suppose that x is real and |x| is large. Then x? is large and v is expected to be similar to the solution of the approximate equation #” + u = 0, In other words, v is expected to behave like sin x or cos x for large valucs of |x{, implying that the solutions of the original DE will behave like sin x/x and cos x/x as |x]—> 0. The conditions under which such statements are true, and the degree. of approxi- mation must, of course, be analyzed and proved (consult hooks and articles on asymptotic behavior of solutions). In this case it may sullice to point out that the 3.6 SOME OTHER METHODS OF SOLUTION 152 exact solutions for v may be taken as vi) = cosx— sinx/x, — va(x) = sin x + cos x/x, and for large |x| (real x), they indeed behave like cos x and sin x. fxample 4. The differential equation y+ Qn+l—xy=0 (=0,1,2,...) is used in quantum mechanics to describe the motion of a harmonic oscillator.* It is known (Example 8, Scction 3.5) that one of the solutions for the case n = 0, the ground state of the harmonic osciltator, reads y(x) = e717, It has been found advantageous to “factor out” e—?7/? for the cases n = 0 as well. Set yO) = eT Fp(x). Then the DE for v(x) reads v” — 2xc’ + 2nv = 0, This is the Hermite DE. ‘The acceptable solutions in quantum mechanics are Hermite polynomials (because » must vanish at infinity). This leads to the solution for y(x) in the form Wx) = Hye", where f(x) is the nth Hermite polynomial. Changes of the independent variable are also useful. Mf, in the DE vy" + P(x’ + Qldy = 0, the change of variable ¢ = f(x) is made, then ~via _ "= eT dt de ae FO 1 FY FV oa, w= FG OP + ELC and the equation becomes 2, L Bryeot + Sur + Peoswl+ avy = 0. If x is expressed in terms of ¢ [by solving ¢ = f(2)], then the DE will have f as the new independent variable. A judicious choice of the function f(x) may simplify the new equation. In particular, if we require that (’() + POYS'G) = 9, which implies that LC) = exp {—J* P@ dé}, then the new DE will have no first-derivative term. Example 5, The Euler differential equation x?y” + xy’ + y = 0 can be simpli- fied by the transformation x = e', f = log x (for x > 0). * Sce Section 11.3. 152 LINEAR DIFFERENTIAL EQUATIONS OF SECOND ORDER The transformed equation reduces to dy a ty=0 and has solutions y,(#) = cos# and yat!) = sins. This leads to the general solution of the Euler DE yx) = c1 cos (log x) + co sin (log x} (as on p. 142). BIBLIOGRAPHY Coppinaron, E. A., An Introduction to Ordinary Differential Equations, Englewood Ciills, N.J.: Prentice-Hall, Inc., 1961. Coopiscton, EB. A., and N. Levinson, Theory of Ordinary Differential Equations, New York: McGraw-Hill Book Co., 1955, Ince, E. L., Ordinary Differential Equations, New York: Dover Publications Inc., 1926. Ramvitze, E. D., Intermediate Course in Differential Equations, New York: John Wiley & Sons, 1943, PROBLEMS 1, Observe that the DE x2y” + xy’ — y = 0 has a solution yi(x) = x. “Use the method of variation of constants to find a second solution and then proceed to solve x?y” 4 xy — y = 1/1 — x). 2. From our experience with the preceding problem, it should not be difficult to find a solution of (1 — x)p"' +- xy’ — y = O by inspection. Using this, solve (ay toy y= . Using the method of variation of constants, show that the general solution of ¥ + w%x = f(1) cam be expressed as ow 2 x) = Acos wt + Bain wt + tf sin oft — r) f(r) ar. B |. Consider the Bessel equation of order zero: xy" + »’ +.xy = 0. Show that the Wronskian of any two solutions is W(x) — C/x where C is a constant, dependent ‘on the particular choice of the solutions. What is the Wronskian of two solutions of the Legendre equation? of the Bessel equation of order #7 ; . Show that the Wronskian of Jg(x) and Yo(x) is 1/x. [Hint: Since W = C/x for all values of x, one may choose an x so small that only the leading terms need to be considered.] What is the Wronskian of Jo(x) and No(x)? of Po(x) and Qo(x)? 1, 6.'Show that the Airy DE y’” — xy = 0 possesses two linearly independent ‘solutions that are analytic at the origin. Evaluate them by the power series method. w x a . Show that the equation x*y” — PROBLEMS 153 |. Solve the DE xy" 4 (<2 4+ x)y’ + y = 0 by the Frobenius method. Can you express the obtained series in a closed form? . Solve the DE xy” + x?y’ — 2y = 0 by the Frobenius method. Show that one of the series terminates. Express the second solution in a closed form (using a method of your choice). Show that the DE (x1 4+ 2x7)y"" + dy’ — 6x¥y_= 0 possesses the following series solutions: viG) = EE Le? + yet = hye + rege =, yo) = O/VE) (+ Be? + Max! — pdegt + ale? - What is the radius of convergence of the series? . Solve completely by the (generalized) Frobenius method the following differential equations: a) xy” + (E+ xdy’ + 2y = 0, by xy” + By +y = 0, ©) x8y” + xy’ + oxy . Show that the DE xy” + {1 — x)y" + Ay = © possesses polynomial solutions if » is a nonnegative integer. Evaluate these polynomials for \ < 4, standardizing them by the condition (0) = 1. . Consider the DE (1 — x2)y" — xy’ + Ay = 0. a) What are the singular points of this equation? b) Solve the equation by the Frobenius method and deduce the values of \ such that one of the solutions will be analytic at the singular points. Show that this solution is a polynomial. ©) Develop explicitly several such polynomials standardizing them by the condition y= 1 |. Show that the DE y” ++ (cosh 2x — 4)y = 0 possesses the solution YO) = x + bx3 — dex? — Agr” — {Hint: Expand cosh 2x in a power series before applying the Frobenius method.] © has no solutions of Frobenius type. Try, however, expansions about the point at infinity and show that ‘ yx) = Ax cosh (I/x) + Bx sinh (1/x). . In the DE y” ++ 2p’ + (1 ~ 2/x2)y = 0, conjecture the behavior of p(x) for large x (see p. 148). Denote this function by y(x) and proceed to solve the DE by sub- stitution y(x) = yaCx)o(). 3. A number of DE's can be reduced to the Bessel equation. For instance, show that the DE y” + x?y = 0 can be reduced to (he Bessel equation of order » = 1/(p + 2) (ee p. 135) by the change of the dependent variable y(x) = /X (2), followed by the change of the independent variable 2 pz rex TAPE 2 +p CHAPTER 4 FOURIER SERIES 4.1 TRIGONOMETRIC SERIES A series of sines and cosines of the type $ + 5 (a.cos nx + by sin nx) “1 is known as a crigonometric series. In most applications the variable x is real. Then sin nx and cos nx are bounded and the series will converge if some rather mitd conditions are imposed on a, and ba. Examples 1. a, = by = Ef/n? (0), ay = 0. The series reads cos x + sinx + }cos2x + sin 2x + boos 3x4 --- and converges absolutely and uniformly for al! (real) values of x (¢.g., by the ratio test and the Weierstrass M-test). 2. a,= 0 (alin), b, = In. The series reads sin x + 4sin 2x + 4sin3x+--- and converges for all x, e.g., by integral test since the integral J” (sin tx/4) dt con- verges for all x. However, the convergence is not absolute; for instance, take the point x = x/2; it is not uniform either (for the entire x-axis). 3. a,=1, =O (allz). ‘The series reads 4+ cosx-+ cos 2x + cosix+-°: and diverges (by the nth-term test) for almost all values of x {the exceptions are points like x = 1/2}. If the trigonometric series converges (whether uniformly or not), then it repre- sents some function f(x), and we can write foy= 2+ F(a, cvs ux + by sin x), pad 154 ¥ 4.2 DEFINITION OF FOURIER SERIES = 155. What kind of functions are representable in such a manner? For a representation by a power series the conditions are (for real x): 1. The function is differentiable any number of times. 2. The remainder in Taylor’s formula approaches zero. ‘These conditions are fairly restrictive. The remarkable property of trigonometric series (discovered by Fourier) is that they can represent functions from a much wider class, including discontinuous functions. ‘One property of trigonometric series should never be lost from sight: By their very nature these series can represent only periodic functions* with period 2x {this need not be the primitive period, that is, (x) may have a smaller period 7, but 2x must be an integral multiple of 7). 4.2 DEFINITION OF FOURIER SERIES Let us assume that a certain function is represented by f®) = z + x (a, cos ax + 5, sin nx), and the series converges uniformly in the interval —x < x > by sin nx cos mx. na a= Sx) cos mx = This series is still uniformly convergent and can be integrated term by term: 4m af cos nx cos mx dx +e te “ f x) 008 mx de = a f cos mx dx + J) —r = ms ° + + Yo by f sin nx cos mx dx, wei "Se This procedure allows the determination of coefficients aq if f(x) is known and is based oh the important property of sines and cosines known as the orthogonality Property: a) [osin nxcosmxdx =0 (all, m> 0), + . wy [esnecormede= [0 nn, —_ ls (ifn = m) 0 (i'n = m) 4 ©) fo sin nx sin mx dx = . -* ns (ifn =m) * So far as the interval (— 00 < x < -|-00) is concerned, af course. 156 FOURIER SERIES 42 These formulas can be derived dircetly, using the well-known properties of sines and cosines. For instance, sin nx sin mx = 40s (n — m)x — 4cos (n + mx. Therefore [0 sin mx sin mx dx = 4 [eos (n = myx dx — af cos + m)x dx. Ifa 4 m, then pte . | cos (1 ~ m)x dx = ifn = m, then * dy = Un. [2 cos — myx dx ~ Also, [7 cosa + mxde= 0 (forall.n, m > 0) and formula (c) is established, Formulas (a) and (b) are proved ina similar fashion. Returning now to our series, we see that all terms in the infinite sums excep! one will vanish. Moreover, [7 cos mx dx = 0 (m> 9) so that [7 10) co8 me de = This relation permits us to calculate any desired coefficient a,, when the function J09) is known. The coefficients & aze treated similarly: The expansion is multiplied by sin mx and is integrated. Orthogonality relations yield ‘ f * 90x) sin mx dx = Finally, to obtain ao, we integrate the expansion fx) = ao/2 + Y (a, cos nx + By sin ax) at as it stands. This results in rr fx) dx = agx. 43 EXAMPLES OF POURIER SERIES 157 Tt follows that the coefficients ay, a,, and &, can be evaluated from the following formufas: tt an Uf" feyeormede (n> 0), di be = (fr) fe {Q)sinnxds (n> 0). The coefficients a, and 6, calculated from a given function f(x) by means of these formulas are known as the Fourier coefficients of f(x). The trigonometric scries constructed by means of these coefficients, 2 4. ¥ (an cosme + by sinns), x is called the Fourier series of f(x). It is important to note that Fourier coefficients can be constructed for a wide class of functions, including some discontinuous functions. 4.3 EXAMPLES OF FOURIER SERIES: Example 1. Consider the function f(x) = x7. Its Fourier coefficients are readily calculated: ay = (fay [x8 ax = Gx, an = /2)f Pcosnx dx = (-1Y in) (n> 0), by = anf x sin uxdx = 0. The Fourier series is easily seen to be uniformly convergent for all values of x and represents a function g(x) = /3 + zy (-1)"(4/n?) cos nx. <1 The graph of g(x) is shown in Fig. 4.1. Itis clear that the Fourier series of f(z) = x? represents 2 periodic extension of the values of f(x) in the interval (—~, +»). atx) Figure 4.4 158 FOURIER SERIES 43 Example 2, Now consider a discontinuous function -1 @<%, 9" lit G20, The Fourier coefficients are 0 - Lf coast [ana -141-0, dg = rf? uf” 4, = if (cos nx) dx + if (+08 nx) dx = 0, edie alo if? ft? bn = 1f (—sin ax) dx + if (4sin nx) dx ade x Jo tt 2 [sinned = [tr = oth ido 0 (a= even), and the Fourier series reads 4a. xy —sin mx. #0) = 5 Zan The series is convergent in the interval (—x, +-) and therefore g(x) is well defined. ‘The graph of g(x) is shown in Fig. 4.2. Explicitly, the Fourier series converges to +1 for 0 zp ' S(xo — 0) = lim f0d, 2 0), and bn $n = arctan 7® (n> 0), then the Fourier series reads 2nnx fey = BF Ay cos GE- )- In many applications, where x represents distance, the period 2£ is most convenient. The formulas read fl) = anf? + SS fy 608 (ex/L) + by sin (nex/E)h, a1 tl ay WL fe ld) €08 (nwx/L) de, i bn 4 WL fe Sx) sin (awx/L) de. : 44 PARITY PROPERTIES. SINE AND COSINE SERIES 16E 4.4 PARITY PROPERTIES. SINE AND COSINE SERIES Suppose that a function f(x) is to be expanded into a Fourier series in the interval (—E, +0). If fQx)is even, then all coefficients 5, must vanish while the coefficients a, are obtained by integrating from 9 to Z only and multiplying the result by 2: u m4 mx, foo = B+ Ze on cos a On, 2f (flo)cos = dx (f(x) = even). Similady, if /(0) is odd, then alt a, are zero and fo = x ba sin" b= 2 [fos as (f@) = odd). These two results give rise to two other types of trigonometric expansions, known as the Fourier cosine series and the Fourier sine series, Suppose that an arbitrary (not necessarily even or odd) function f(x) is given. If we calculate the “hall-range integrals” and define L 2 nex = i f(x) cos | dx, z+ FE aycos aaa and form the series then this series will reproduce the function ff O and the Fourier sine series reads gw) = 3 > dein 25 — 1 > | sin HEE. * a m a Finally, the Fourier cosine (half-range) coefficients are 1 3 2 x 1 tex cos aw — 1 40 = 5° an? lf (+ feos ae = Seer, and the Fourier cosine series reads a@=i-32 o 1 cos 27% ® Lb m All three series are shown in Fig. 4.4. ne — * [ . 7 . : a) ee = a a ee Bl) 4 Me x Figure 4.4 The formulas for the Fourier sinc and cosine series for “half the range” are related to the fact that the functions sin nx are orthogonal over the interval (0, +): [sia nx sin mx dx = 0 (a # m), 0 and the same is true for the functions cos nx: ff cos mx cos mx dx = 0 (a x m). fo 164 FOURIER SERIES 44 However, they are not orthogonal to each other, namely, the integral f sin nx cos mx dx 0 is not necessarily zero. There is a remarkable proof of the orthogonality of sines based on the knowl- edge of two facts: 1. The functions sin xx satisfy the DE y” 4+ #?y 2. The functions sin nx vanish at x = O and x = 7. The method is as follows: Write Ye = sin nx, Yn = Sinmx {0 * m). Then Yt wy, = 0, Wh + myn = 0. Multiply the first equation by y,,, multiply the second one by y,, subtract, and integrate from x = Oto x= «2 fF om — vibra + 0? — m)f” rntm doe = 0. In the first integral, use integration by parts: ” yom dx — [" yilan dx = vhym|” ~ [vive de ff vim de ~ [vida dx = vas” ~ fy ~ vhanal + [plot de = Ohm — ade) ote o Since yp and y, vanish at x = O and x = 7, it follows that this expression is zero. Therefore @- mB” Ym dx =0. 0 Since n > m, it follows that f£ Yaya ax = 0. 0 A similar proof can be developed for the cosines: The function cos 7x has the following properties: 1, It satisfies the DE y” + ny = 0. 2, Its derivative vanishes at x = Qand x = 2, The proof is essentially the same except that now the expression 4m — varndle i vanishes because the derivatives y;, and y;, vanish, Finally, consider a similar proof for both sines and cosines over-the interval (—*, +). Let p, denote either sin x or cos nx. 45 COMPLEX FORM OF FOURIER SERIES = 165 We make use of two facts: I. yp satisfies the DE yf +n? yy = 0. 2. yn is periodic: yu(x -+ 2m) = n(x). In this case, the integration is from —x to +x, and the statement Ohm — Herne = 0 is made on the basis of periodicity. The result, namely Yond = 0 (forn ¥ m), establishes, at once, the three orthogonality relations: [sin axsinmxdx=0 (n¥ mi), [U cosmxcos mx dx ~ 0 (a x m), cr sin axcos mx dx = 0 (nx m). Observe, however, that the method employed does not prove the orthogonality of sin nx and cos nx, namely, +r f sin nx cos nx dx = which must be established separately. 4.5 COMPLEX FORM OF FOURIER SERIES The Fourier expansion fo = F+D ( cos “ty sin ™) (-Léx 0), Cn = {Rta + in) (a < 0), hag (a = 0), 166 FOURIER SERIES 45 Then the Fourier series can be written in its complex form: te Say= Eee (LS x SL). The convenience of this form is obvious. From the formulas for a, and }, follows the formula for ¢,: Exercise. Deduce this statement by a detailed argument. Alternatively, the above formula may be derived by multiplying the series by eit!) aod integrating. The complex exponentials are readily shown to be orthogonal in the senset that auf? @*m), +E dlarett) stat) f € e iE aE (n= m), and the formula for cy follows. Remark, Although the Fourier series now appears in complex form, its sum f(x) is still assumed to be real. In this case the foliowing properties arc easily verified: 4. €9 is real; c_n = cf, 2. If f(x) is even, all c, are real, 3. If f(x) ix odd, cg = 0 and alll ¢, are pure imaginary. Now consider the complex functions of the real variable x. They can also be ex- panded into Fourier series and the complex form of the series is now a natural one. The formula for c. docs not change, but the above three properties of the coefficients no longer hold. ‘One may show, however, that a) if f(x) is even, then c_, = cq, and b) iff() is odd, then ¢_4 = —eq (and co = 0). Example. The function lO (-r For all values of x such that —x < x < x, the series uniformly converges to f(x) in any interval x1 —m, x2 (Ancos kx -+- Be sin kx), ai where the coefficients A; (k = 0,1,...,) and By(k = 1,2,3,...,#) are undetermined. Let us define Ay and By by requiring that the total square deviation + Da =f" 92) — end dx be minimized. Straightforward calculation yields . te +e Dy =} OP de + [diese —Aof fe) ay| » +. + DO jedi - aa f Sx) cos kx dx} mi ~ » + + epi ~ am see) sin kx a} a a Exercise. Using the orthogonality relations, give all (he details leading to this formula. The whole expression is minimized if each term in { } is minimized, For instance, the quantity dy = wAE = 24e{"” fla) cos kx dx (k= 1,2...54) has an extremum if + ge = dedy— 2) flx)coskxde = 0, or Lf” i Aas Fcoskxdx (Kk = 1,2,...,0) — This extremum is a minimum since d?4,/dA2 = 24 > 0. In this fashion we also obtain B= (sy f fay sin kx dx (= 1,2...57) Ay = anf Sx) dx. It is seen that the total square deviation is minimized if A, and B, ate chosen as the Fourier coefficients of the function f(x). 47 CONVERGENCE IN THE MEAN 17k Now substitute these optimal values (denoting them as a, and },) of the coefficients into the expression for D, to obtain the minimum total square deviation: +" =| Voor ax — [sro + «cok + 8} [Dyh Since D, cannot be negative, the following inequality holds: +e aha + Sak + 8) (cos nx/n”). = Can you justify the validity of term-by-term integration ? c) Set x = 0, and show that 2_1 (1/n?) = 2/6, d) Develop this idea further to establish the result XY d/nt) = x4/90. a e) Verify your result in (d) by means of Parseval’s equation applied to the function ‘folx) obtained in (b). . A damped harmonic oscillator under the influence of an external periodic force £(0) obeys the differential equation. mE +E + RX = CN. Assurting @ steady-state solution, salve the problem by the Fourier series method. In particular, develop the explicit series for x(t) if f(x) is the triangular wave of Problem 4 above, . A simply supported beam like that shown in Fig. 4.8 is loaded by a variable joad (per unit Jength) q(x) = (@/L)x. a) Show that the deflection is given by dott 2 (1 ax yx) = aH ee b) Solve the problem in closed form (integrating the DE) and show that the error for the deflection at the midpoint is only 0.387%, provided only the first Fourier term is used. 178 FOURIER SERIES 14. A periodic (but not harmonic} electromotive force E(t + 41/7) (-T/2 <1 <0), Eo(l — 4t/T) (<1 0). Observe that if s is complex and s = o + iw, then FG) = f "tele dt -f te cos wt dt — if te~** sin wt dt 0 oO Both integrals converge if ¢ > 0 and diverge if o < 0. For o > 0, the actual evaluation is conveniently done in the complex form (using integration by parts). 2 fO=1, Fe) -f ev"'dt = 1/s (Res > 0), Using integration by parts and induction, 3. ft) = 1" (n= integer), FCs) =f te-"dt = nl/s"*! (Res > 0). Using the definition of gamma function, 4 fO= @>-D, FG)=TEtY (Res > oy. sr 1 5. f= 6", FQ) = rag (Res > a), 6. A= sink, FU) = gas (Res > 0), 7. ft) = coskt, Fs) = a (Res > 0). From these examples, it is evident that the Laplace integral converges, as a tule, for a restricted region of s (in the complex s-plane). Jt is general feature of the Laplace integral that this region can be characterized by Res > a, where ais some real constant. In other words, the Laplace integra! converges to the right of some vertical line in the s-plane (Fig. 5.1). Line of convergence Region of divergence fea Figure 5.1 182° THE LAPLACE TRANSFORMATION 5.2 OF course, there are functions for which the Laplace integral diverges for all values of s; in other words, these functiogs do not possess a Laplace transform. For instance, f(r) = e!”. The integral fy e!~*' dt diverges for all (compiex) & Thus the first problem in the theory of Laplace transforms is to determine the class of functions and the values of s for which the transform exists. Definition 1. A function defined on the (infinite) interval a <¢< 0 is called piecewise continuous on (a, 0) if for every finite interval a <1 0. Existence theorem. If (6) is piecewise continuous for 0 < ¢ < oo and is of exponential order ao, then the Laplace integral Foy = £0} =f" "sod converges for Res > og. Moreover, the integral is absojutely and uniformly convergent for Res > a; (7, = real), where go, > oo. Proof. Lets = 0 + iw and consider Gals) =f" er fen| ae — f" le" iso at 0 lo There is a finite number of discontinuities (say, m) within 0 < ¢ < R; denote them by f1, f2,..., fx, and, furthermore, let t9 = 0. Then ‘pk im optins R ox =f error = & [enon s [eryora m tig 0. Since it is monotone nondecreasing (a nonnegative integrand), it must approach a limit: ’ firm Gr(s) = im "te PUI ae =f 4 | dt = Gs). The existence of G(s) implies absolute convergence of the Laplace integral and, therefore, the existence of the Laplace transform 2yen) =f esd Res > 00) ‘The uniform convergence for Res > o, (with o1 > oo) is established by the Weierstrass M-test for the integrals: f ef] dt < f le—')| dt (for all R). Observe that the Laplace integral is convergent in the open region Res > oo, but uniformly convergent in the closed region Res > 04 (where a1 > 9). Remark. The property of piecewise continuity ensures the existence of finite integrals {extensions are possible); the property of exponential order ensures the convergence at infinity. The conditions on f(t) prescribed in the existence theorem are sufficient but not necessary. There exist functions which are not of exponential order but, never- theless, possess the Laplace transform, for example, S(t) = ie” cose”, This function is not of exponential order because e” > Me for sufficiently large 4, no matter how large M and ¢ may be. However, the Laplace integral f ” e*21e" cos ef dt 0 does exist for Re s > 0. Indeed, integrating by parts, we have f 7" 2te" cose” = e sin fl + sf etsine dt 0 oo = —sin t+ sf” ec sine” dt. 0 Since [sin e*"| is bounded the fast integral exists for Res > 0. ‘Therefore f(#) possesses a Laplace transforni. Jt is fair to say, however, that functions like this are extremely unlikely to appear in physical problems. [84 | THE LAPLACE TRANSFORMATION 53 $.3 BASIC PROPERTIES OF LAPLACE TRANSFORM ‘The fundamental feature of the Laplace transform, implied by the very nature of the Laplace integral, is that two functions which are identical in the range 0 << o but different otherwise have the same Laplace transform, For instance, the step function St) = ? «<9, 1 @>9 has the Laplace transform LS} =f edt =1/s (Res > 0), which is identical with the Laplace transform of the function f(#) = | (all 4). In general, the two functions (see Fig. 5.2) fi(t) = f(t) and fo(2) = MSW will possess the same Laplace transform. 0 ASK) *t +t Figure 52 Remark. A similar situation occurs in Fourier series: Two functions, fi(2) and fox), which are identicat in the interval —x < x < x but different otherwise will possess the same Fourier series (of period 2m). As a matter of fact, if a certain function f(r) has Laplace transform Fs), it is more convenient to treat F(s) a8 the Laplace transform of f(t)S(2) rather than of f(t), The reason for this may become clear from the following example: Suppose that the function f(S() is “shifted” by a length @ fo the right (Fig. 5.3). The Laplace transform of f(t — a)S(t — a) is LYU — aS(t — a)} =f et — aS — a)dt (a > 0). We make a change of variable, ¢ = + + a, and obtain £ em" ft — a)S(t — a) dt = f eA fr )S(r) dr = " f * e(r)S(r) dr, or £{fC — aS — a} = e PLY} = @ > 0). This relation is often called the shifting property (or the translation property): If a function f(s) is “chopped” by multiplication by 5(s) and the resulting function SSCA) is translated to the right by an amount a, then the Laplace transform is 53 BASIC PROPERTIES OF LAPLACE TRANSFORM 185 Original f(s) for £<0 _-| S50) pe aa t 2 +f Figure §.3 multiplied by e-*'. The reference to “chopping” is then avoided if the above formula is rewritten in the form EYE — QS — B= MLNS (> O- It is worthwhile to emphasize the condition a > 0. The shifting property does not hold for a <0. Whether or not the function (1) is “chopped,” if it is translated to the left, the new transform bears, in general, no relationship to the original one. ‘Armed with the knowledge of shifting property and observing that eth = Vs, ete} = Ve-@ (Res > a, it seems logical to inquire whether an analogous property will exist if f(0) is multiplied by an exponential function. Indeed, ele poy = [PMO at = fc rod 0 0 Setting s + a = 7, we obtain f eH) dt = Fi) = Hs + 4) so that Le f(N} = Fs + a. where Fs} = £{f(0}. This result may be called the attenuation property (or the substitution property): IE f(0 is “attenuated” by the exponential factor et, then the transform is shifted (to the left!) with respect to the variable s. Observe that the attenuation property is valid whether a is positive or negative. However, the transform usually has a different region of existence [as Res > a for I/{s— 2) yersus Re s > 0 for I/s}. Remark. The term “attenuation,” borrowed from electrical engineering, ordinarily means multiplication by et, where a > 0. This is the case in most practical applications, and therefore the name attenuation property is more descriptive than the name substitution property, implying that s + @ is substituted for s. ‘The most important property of the Laplacé transform is, perhaps, the simple relationship between £{/()} and SUC). Suppose that /(4) is continuous, 186 THE LAPLACE TRANSFORMATION 5.3 and integrate the Laplace integral by parts: - at = se |? ites fre (i) dt = — (1/s}e rol + lsfre P(t) dt or sf * efi) dt = f{0) + f£ ef) at. 0 fo Assuming that 2{/"(1)} exists, this may be written as £{f'(O} = sk fO} — /@ and is known as the derivative property. Remark. ntegration by parts is not permissible unless f(¢) is continuous. On the other hand, f"(t) may be piecewise continuous. For an extension of this formula when f(s) is piecewise continuous, see Section 5.7. The derivative property forms the basis for most applications of the Laplace transform and sheds tight on Heaviside's operational calculus. Example. A particular integral of the DE y’ — y = e7* is sought. Multiply the equation by e~** and integrate from 0 to a. Let £4y(x)} = YO). Then, employing the derivative property, deduce SYS) — yO) — YO) = 1s + 1). Suppose that (0) = 0. If an arbitcary particular integral is sought, this choice is just a matter of convenience. Then (s — 1)¥(s) = I/(s + 1), and —— tt dt di, G4D@—-D-2s-1 2sFi It is known from the preceding examples that ' ig{e} <1/(s—D, and gfe} = 1/8 + D. It follows that ¥(s) is the Laplace transform of the function y(x) = 4e* — $77, which is, indeed, a particular integral of the given DE. Tt is not difficult now to derive a formula for the Laplace transform of the second derivative and, for that matter, of the derivative of any order. Replace IW by f'(2 and J’) by £’"() in the derivative property and obtain : LEO} = 2} - £'O - LEO} = s72(FO} — 5/O) — £'O.- ¥(s) = or 5.4 ‘THE INVERSION PROBLEM 187 This formula is valid, of course, provided f‘() is continuous s (and ff, and f" have a Laplace transform). The general formula for the ath derivative, established by induction, reads 2) = ste} — EY, mi where f(r) is the mth derivative of f(1) and f(0) is its value at ¢ = 0. 5.4 THE INVERSION PROBLEM The example of the preceding section involving a DE provides a blueprint for the use of the Laplace transform in solving similar problems: The relation satisfied by the unknown function f(£) is subjected to the Laplace transform. The result is a relation satisfied by F(s) = £{f(0}. From this relation F(s) can (in principle) be determined. ‘The third step is to find the unknown function /() from its Laplace transform F(s). This last problem involves performing the so-called inverse Laplace transformation. Relation satisfied Relation yt BH? {hopefully si it Step 1: Transform: T (Direct solution difficult) Step 2: Solve for F(s). i found : A os ‘oe Sten 3: Invert Figure 5.4 ‘The entire idea is very conveniently expressed in the following “flow diagram” (Fig. 5.4). The first step in this scheme is straightforward: Laplace integrals have to be evaluated. The second step, finding F(s), should be rather simple {if it is not, then there is little point in using the Laplace transform at all). The third step is, perhaps, the most difficult one and a variety of methods is used to perform it. The inversion problem is greatly facilitated by the fact that the Laplace trans- formation represents a litear operation, that is, it has the following two funda- mental properties: a) £{af)} ~ ae{f@)} — (@ = const), b) LAO) + AO} = LENO} + LAO}. ‘These two properties stem from the fact that the operator & which transforms SG) into F(s) is an integral operator.* * Properties (a) and (b) are general properties of integrals. 188 THE LAPLACE TRANSFORMATION 5.4 Bearing this in mind, we expect that if a transform F(s) can be split into two parts, FG) = Fy(s) + Fas), then it can be inverted term by term, writing SO = AO + AO, Fis) = £{f(0}, and Ff) = £(/20)}. Symbolically, this is written as © MO + Fy9) = 27H) + HAO}, where the inverse operator &~* transforms the function F(s) into the corresponding function f(t). Similarly, we expect that where £7 'aF(s)} = ak-{F{s)} (a = const). These relations mean that £—? is, presumably, also 2 linear operator. , A Figure 5.8 Remark, The linearity of £ alone is not sufficient fo prove that £~! js linear as well. What it actually implies is that the two functions FU) = &-HF1G) + Fo} and ft) = 2° UF} + 27HFoG)} have the same Laplace transform. However, does it mean that f(#) and 7(¢) are the same? Not necessarily so, even if only the values for 0 <4 < © are taken into consideration, For instance, the two functions (Fig. 5.5) A@=1 @ir> 0, 10 (¢ = integer), Alt) = “ I(t = not an intezer) + ; are not identical but have the same Laplace transform. The difficulty would disappear if the inverse transform were unique (no two different functions can have the same La- place transform). 55 THE RATIONAL: FRACTION DECOMPOSITION 189. From a certain point of view, c.g., a physicist’s, functions tike f,(#) are “ab- normal’ and should not be taken into consideration. Then, perhaps, f,(f} can be regarded as the inverse transform of F(s) = 1/s. To put this idea on a firm basis, various theorems have been proposed. For instance, it can be shown* that if we confine ourselves to (real) functions which are a) piecewise continuous and piecewise smooth (i.c., have a piecewise continuous derivative) on (0, 00), b} of exponential order, ¢) defined by 3 /Gxo + 0) + f(x — 0)] at each jump discontinuity xo, then two such functions possessing the same Laplace transform F(s) must be identi- cal. In other words, the inverse transform is unique within the class of such functions. This theorem is adequate for a great majority of physical problems. 5.53 THE RATIONAL FRACTION DECOMPOSITION Jn a large number of practical problems, the function F(s) to be “inverted”} is a rational function of s. In this case, a very general and efficient method can be developed for the solution of the problem, as illustrated below by a series of examples. Example 1, Find /(0) if Fis) = 14s? — 5s + 6). It is easily checked that this function can be decomposed into elementary rational fractions: FG) = 1 G-H—37~ s In the spirit of the discussion of the preceding section, we can conclude that go = otf} ot {to} - a at ‘The idea used above can be applied to all rational functions F{s), namely, functions of the type F(s) = P(s)/Q(s), where P(s) and Q(s) are polynomials in s. Without loss of generality, we may assume that Q(s) is of higher degree than P(s). Indeed, if this were not so, then we could divide P(s) by Q(s) until a remainder of the desired form is obtained, namely, F(s) = R(s) + P(s)/ 2), where R(s) and P,(s) are polynomials and Q(s) is of higher degree than P,(s). * See, e.g., Churchill, Operational Mathematics, Chapter 6. + That is, subjected to the inverse Laplace transformation. 190 THE LAPLACE TRANSFORMATION 5.5 Remark. The case deg P(s) > deg O(s) is, actually, not likely to occur in practice since @ polynomial R(s) does not possess an inverse Laplace transform.* The decomposition of a rational function is based on the knowledge of roots of the denominator Q(s). If the coefficients of Q(s) are real, then it is known from algebra that Q(s) can be factorized (apart from a constant factor) into factors of two types: 1. (@ — 7)" for each real root 7 of multiplicity m; 2. (5? + as -+ 5)" for each pair of complex conjugate roots (satisfying s? + as + 5 = 0) of multipticity n. Corresponding to each real root r there is a set of sm terms in the decomposition of P(s)/Q(s), of the type At 4 Ae og As G=Ht GAT GAT TG Corresponding to each pair of complex conjugate roots, there is a set of terms of the type Bist Cy _ Bas + Ca eee gp Past Co @ + as + 5) Git as F op @ + as + by Each of these elementary fractions can be inverted in a straightforward manner, using the attenuation rule and the transforms of the powers of f and trigonometric functions. In particular, + (A; = const). (B,C; = const). = leSal-4 “4G ean" by the result of Example 3 of Section 5.2 and the attenuation property. For quadratic factors it is convenient to complete the square: s? + as +b = (s + a/2)? + (6 — 47/4) = (8 + /2)? + 7. {Note that 6 — a?/4 > 0; otherwise, the roots would not be complex. Therefore c is real.} Then we can write Bs+C_ | Bs + a/2) + (C — (0/2)B) Fats (+ af}? Fe sta/2 + 16 Je — @/2)4]_ Bot afy taterapyra From the results of Examples 6 and 7 of Section 5.2 and the attenuation rule, we can deduce that : ' 1 (+ 0/2) | _ paix £ {a Gt afiy +l = Be cos ct * See, however, the discussion of the delta function in Chapter 6, in particular, p. 232. 5.5 THE RATIONAL FRACTION DECOMPOSITION 191 sin cf. and ge | C= @/De | Fafa + FF A general inversion formula for Bs + 4/2) + {(C ~ (o/2)B] (G+ of2 + C= G/B cian ¢ : > t), can also be derived. However, in the case of multiple complex roots it is probably more convenient to employ the general complex technique used when Q(s) has complex coefficients. In this case there is no need to introduce factors of type 2 in the factorization of O(s): The polynomial G(s} can be represented as a product of factors of type | with complex roots, and the inversion can be carried out in complex form. Jf necessary, the final result f(f) can be split into real and imaginary parts at the very end. Exercise. Show that the attenuation rule is valid even if a in F(s + @) is complex (this is necessary for inversion in the complex form described above). Example 2. Find f(z) if _ 3 -st 16 *O = GI GF DEO) The factorization of the denominator is already performed and the decomposi- tion is of the form Fo) = 5 Therefore ACs + 2N6 ~ 6) + Bs — IMs — 6) + Cs — INS + 2 @-— DE + DG — 6) Is 416 This implies the identity that is vatid for al? s: ACs + M5 — 6) + BG — INS — 6) + C(s — IN + 2) = 3s? — 25 + 16. In principle, we could obtain A, B, and C by equating the coefficients of powers of 5. However, this is not the quickest method:* It is more profitable to observe that since the above equation holds for all values of s, we may set s equal to any number we please. If we set s = 1, we immediately obtain AL DU — 6-3-2416 of A= —]E + It leads to three equalions with thie unknowns. For more complicated rational functions, this method is quite cumbersome. 192 THE LAPLACE TRANSFORMATION 55 Similarly, with s = —2, we obtain B = §, and with s = 6, we obtain C = 33. ‘The remainder of the problem (inversion) is straightforward: S(O) = Hel + $e + Het It is not difficult te see that this method will produce quick results whenever Q(s) has simple roots. In fact, this situation can be disposed of once and for all by the following formutas: Q0) =A ii (s—ri) (A = const), «) where the symbol J] designates the product of 1 factors, TL 6-1) = 6 — 6-2) = tet with ri = 1, 2, 3, .. 2} being the 2 (simple) roots of O(s); FO) ~ POY QU) = WAS ills ~ ro @ 4p = PEM Ts) GL) oO) where the symbol n 0 ia Means that the product is taken from # = t to i = », but with the factor corre- sponding toi = j omitted. Furthermore, f= UAE ae. ro) This extremely helpful result is sometimes called the Heaviside expansion theorem. An alternative form of this theorem reads: If all roots 7; of Q(s) are simple, and if the degree of P(s) is less than the dearer of Q(s), then oo {PO ye Pen et, Qs} z o@)* where Q’(r;) is the derivative of Q(s) evaluated at s = r;. This form of the theorem follows from the fact that if 06) = 4 Ts = 12, then ns i oa =4Te- 1 inf (as verified by elementary calculus). 55 THE RATIONAL FRACTION DECOMPOSITION 193 Example 3. Find f(0) if F{s) = 1/68? + s?+ 54 0. Although Q(s) is not written in factorized form, our knowledge of algebra should tell us that if Q(s) is multiplied by s — 1, then b-N@O+s? 54+) = 5-1 Therefore s s-1 t FO) = se @FDETIE- DN" PNET Although the real decomposition Bs+C roy = 44 Bete is available, we shall employ the complex decomposition F(s) = A stitsgits By using the Heaviside expansion theorem, we find that ~ CHEN Hat td 1 1 oe TH acl c= so that fo= dete UPD ey Clad This expression is easily brought to real form: ; f(t) = fe — boost + fine. Example 4. Find /(), given that F(s) = s?/(s ~ 1)°. According to the rules of algebra, F(s) can be decomposed into elementary fractions: B "= Gon t ot a Then _ A — 1? + BE -Y+C_ Fe) ~ G-1F -C-» implies that As — 1)? + Bs — + C= 5. 194 THE LAPLACE TRANSFORMATION 5.6 According to the idea of Example 2, we can obtain C easily by setting s = [; then C = |. However, any other value of s will yield a system of equations for A and B. Can we obtain A and B by an easier method? Our experience in calculus may suggest that we differentiate our relation with respect to s: If it holds for all s, then 2A(s — 1) + B= 2s must also hold for all s. If we set s = 1 now we immediately obtain B — 2. In the same spirit, we can differentiate once more and obtain 24 = 2,50 that A = 1. ‘Our result is then 2° 1 2 1 FO> GTR Gat GF HS so that SQ) = ef + ete + eth? = eft + 2¢ + 67/2), The idea employed in this example may be generalized and expressed in terms of the following theorem: General Heaviside theorem. Let F(s) = P(s)/Q(s), where P(s) and Q(s) are polynomials and the degree of O(s) is greater than the degree of P(s), and let s =r be a root of Q(s) of multiplicity m(m > 1). Then F(s) can be de- composed as follows: F(s) = H(s) + 5: 4 soa (Ss — 7k where H(s) is finite at 5 = r (it may be zero) and 1 ant Ae = Ga I dg OO), with G(s) defined by G(s) = Fists — J". Observe the similarity of the expression for A, with the formula for the residue (if k = 1) of a function having a pole of order m. Exercise. Prove the general Heaviside theorem. [Hint: Consider F(s) as a function of the complex variable s. Show that it can have only isotated singularities and consider a Laurent expansion about one of them.) 5.6 THE CONVOLUTION THEOREM Consider the problem of forced oscillations of a (damped) harmonic oscillator. It is described by the DE KA DME+ whe = f(t) (= dx/dt, ete.) 5.6 THE CONVOLUTION THEOREM 195 and initiat conditions such as x(0} = x9 and %(0) = v9. The Laplace transform method calls for taking the transform of the DE, S7X(s) — sxq — Vo + 2AsX(s) ~ Bx + wEX(s) = Fils), where F(s) = 2{f(O} is known, and X(s) = ©{x(f} is sought. This equation is now solved for X(s): Axo + Yo + 5x0 FG) As) = 2 2 2 2 F+Ds+aq s+ 2s + wo = Xs) + Xa(s). The first term may be inverted once and for all. It is logical to expect, and easy that it gives rise to the solution of the homogeneous equation with the given initial conditions x(0) = xo, ¥(0) = vp. Indeed, use the technique of rational fractions: 2dxXq + By + x9 _ Xo(s + A) . Do + Xo FEM + ok (EAM +GR— 2) (HAP 4 WY) tacitly assuming, just to be specific, that a3 > 7 so that the roots are complex; this is the more common case in practice. The right-hand side of this equation is the transform of xi() = xoe™ cos wt + Pot Meee sinwt (w= Va — 2%), and this function does indéed satisfy the homogeneous equation and the stated initial conditions, The entire problem, then, reduces to the inversion of xy) = =O. s+ Das + od We have seen that the function x(t) = £—'{X9(s)} represents the response of the harmonic oscillator to the external influence characterized by the function F(#) {or its transform F(s)]. Moreover, the function xo{r) is expected to satisfy the initial conditions x2(0) = 0, %2(0) = 0, in order that the entire solution will satisfy the given initial conditions. In other words, it represents the response to the function f(¢} of a harmonic oscillator initially af rest. From the poiat of view of the general theory of differential equations, the function xa(z) is then a particular integral of the nonhomogencous equation, while x1(f) is a complementary function. Remark. A subtle difference between the decomposition of x{r) into two parts in the Laplace transform method and the complementary function method is that in the former case both parts have specified initial conditions. In the latter case this is not so and the adjustment of arbitrary constants is made at the very end. 196 THE LAPLACE TRANSFORMATION 36 Now we turn to the general problem of inverting F(s) X,(s) = -;———_,' ee SE DASE The structure of X;(s) is always the same: It is a product of two functions of s, namely, 1 do GS) = ————;' FG) am © s+ das + al The inverses of these two functions are known: One is f(s) and the other is 1 a = 2" basal = tem sinat = = Vu2 — 0) Remark. The function on the right-hand side vanishes at ¢ = 0, but its derivative does not. ‘This may seem like a contradiction of a previous statement if we set F(s) = 1, However, such a conclusion is premature because F(s) is not an arbitrary function of s, but rather a transform of some f(s). This actually rules out the possibility F(s) = 1 because a function such that £{/(O} = 1 does not exist ‘The fundamental problem now is to find 2 (F(s}G(s)} if £7 {F()} = SO) and £7 {G(s} = g(t) are known. The answer lies in the so-called convolution theorem.$ Definition. A convolution of the two functions f(t) and g(t) is denoted by (f+ g) and is defined by (Fes) =f' seat — 7) ar. Note that (f+ g) = (g* f) which is easily verified by the change of variable n= f—7, The geometrical illustration of this concept is as foflows: Suppose that two functions f(r) and g(7) are given (Fig. 5.6a). The curve g(7) is imagined alt-) Figure 5.6 } See, however, p. 232. { Or Faltung theorem, based on the German word faltung, meaning “folding” in Engtish. 56 THE CONVOLUTION THEOREM 197 to be drawn on a separate sheet of paper which is folded to the left, along the line 7 = t/2 (Fig. 5.6b) yielding the curve g(f — r) for 0 < 7 < 1/2. Similar folding to the right will produce g(¢ — 1) for (/2 < 7 < ¢ (Fig. 5.7). A better description of this process would be, perhaps, a reflection of g(r) with respect to the liner = 1/2. Figure 5.7 Theorem 1. If f(#) and g(t) are piecewise continuous and of exponential orders « and 8, then their convolution (/ + g) is continuous and of exponential order ¥ = max (a, 8) + ¢, where € is positive but arbitrarily small, Proof. Evidently, f(r )g(t — 7) is piecewise continuous for 0 < r < 1. Therefore, the convolution integral fy f(7)g(¢ ~ 7) dr is a continuous Function of t.$ More- over, from US Me™ dar) S Mae”), it follows that If S(r)gt — 1) ar| 8, then eft eu ht so that {as desired}. Ia < @, then ‘ I ad 7. so that faa $5 Mills $1 (as desired), Finally, if « = 8, then [ff fede — ny dr] < oye fo t Consult the theory of the Riemann integral in texts on advanced calculus. 198 THE LAPLACE TRANSFORMATION 5.6 r 1 Figure 5.8 Figure 5.9 However, t < Mget! for arbitrary « > 0 and some M3 (depending on ¢, but not on 1). Therefore IU * al < MiMaM get! (as desired). Theorem 2 (Convolution Theorem). If f(t) and g(1) are piecewise continuous and of exponential orders « and 8, then LY +9) = FOG) — [Res > max (a, 6) where F(s) = £{f(0} and G(s) = efg(o}. Proof. By Theorem 1, 2{(f+ g)} exists and is defined by 2{(fe 8) = lim f et auf FOU — dr = tie I Consider the (finite) integral fp and transform it as follows: Ip is a double integral over the shaded triangle in Fig. 5.8. Interchange the order of integration: Rh in . _ fk a Th =f arf eo" fa)ols ade =f Ser) arf eet — x) dt. Perform the change of variable ¢ - 7 = u, dé = du, to show that =f” Ro ace tr) In= f fayarf e g(t) du. The region of integration in the wr-plane is shown in Fig, 5.9(a). Split this region into three parts as in Fig. 5.9(b) and write REZ REZ i Th = Ta + Ins + Ina = f o'r) dr f eM g(u) du ' +f BPP 8405) dr fr e-™ g(a) du + [erm if. pc) du, 5.6 THE CONVOLUTION THEOREM 199 Let R > » and observe that lim fav = tim? ofr) dr tim [Pe et) da = FOGLE). Row Ri2—0 JO Rj2-+0 JO It remains for us to show that limase fez = and lime Teg = 0. Indeed, RAZ Hor Meal $f" "ode f Meta) de 5 ie RID nar Roam Sf eae fe" letao] du. 5 a2 Since the integral ft oe" le@)| du 0 converges (by the existence theorem), the Cauchy criterion for inyproper integrals implies that fe e“lewl du 0) Riz for R = sufficiently large, and Rn nal < of °° e“UfC)) dr. a Since the integral on the right-hand side converges, it is bounded, fovea plat ied: where F(s) = £(/(0}. This theorem foltows readily from the convolution theorem by treating 1/s as G(s). Exaniple. Find f(#) if Fs) = 1/(s? + 45). 57 ADDITIONAL PROPERTIES OF LAPLACE TRANSFORM — 201 Instead of using the rational fraction decomposition, we can write Fs) = (/sMI/G + 9] and deduce that “ fa -f vd = Lae™. If the Laplace transform of a general indefinite integral of f(t), sO=f'faldr @>a>, le is desired and written as G(s) = £{g(t)}, then integrating by parts, we obtain f oR dt = esto! +5 f e"*g() dt. fo ods If g() is of exponential order, then for Res > # (some A), tim e-"g(t) = 0. F(s) = —g(0) + 366s), Therefore or ais) = FO! f fd W>a>0). It now remains for us to inquire when g(s) is of exponential order. If f(A) is of exponential order a, then f "ft ar so that g(t) is of exponential order a if a > 0, and of exponential order zero if « < 0, and of exponential order € (any ¢ > 0) if a = 0. le@| = t suf dr= Mo, a Example. Suppose that we wish to find the Laplace transform of ‘ Wa t 2 f _a 2Wa f mest = a = do. gi) = erf (v/a) Vile © du Va de® Asa first step, we find, by direct integration, the transform of f(#) = e*"(@ > 0): -at™) _ sal? ef plat?) ate yn free a= fre dt, In cases like this the technique of completing the square is found to be convenient: at? + st — (a + s/2Vay — 3?/4a. 202 THE LAPLACE TRANSFORMATION 5.7 Therefore afer} = ete Lc eV RVa ay 0 Now, set (n/a + s/2\/a) — w, and obtain gfe} = ote 1 om’ de = ME 4 erfe va raya” Wat Wi , Since t ast ve eo" dr = “af eva), I away the primitive function theorem can be applied and we can deduce that eter ova) = N82 if oa] = 2% wat ee") so that 1 s {ert (o/a 46 orfe S_. fort ea) = pM ete Fe ‘a Sometimes an occasion arises for us to employ the differentiation or integra- tion of transforms: The expression Fis) = f “eo f(i) dt is a uniformly convergent integrat for a suitable range of values of 5 (see the existence theorem). Therefore oF) [ end or E= — sep}. This formula can be generalized: , ar to feh Fe = (I ee yo}. Let us suppose now that Fs) = £{/(0)} and we wish to integrate F(s). In particular, consider G(s) = f ” Fe) do. By definition, G(s) = f * do f a F(t) de. “Io 57 ADDITIONAL PROPERTIES OF LAPLACE TRANSFORM 203 By uniform convergence, the order of integration can be interchanged and = ° al at Gs) -[ KO af eds =f foe at. In other words, [ Flo) do = & {na}. Example. Evaluate £{sin w/t} and £{Si (¢)}. We know that £{sin wt} = w/(s? + w%. Therefore e ¢| -f aPae arctan 7 x 5 s 5 — arctan = = arecot~ + 2 @ o ls t Also, since si -f ee du, it follows that . L, Ising 1 £{Si ()} = 5% fins = 3 arecot s As the last topic of this section, consider the extension of the theorem on the transform of the derivative for the case of discontinuous functions. Suppose that f(f) is continuous for 0 < 4 < oo except for a jump discon- tinuity at? = 4). Then /’(£) is not defined at all at * = ¢,, but we may define the Laplace integral of f(t) by means of the following equation: f, “et dt = fim f “ee dt + tim f oP) dt. ° wen? mea (We have tacitly assumed that /‘(¢) is piecewise continuous.) Now the integration by parts can be justified for cach integral separately and we have ' £(f'O) = lim e=“f(O|" + tim “fol? +f "eH dt, sh are oo we we where the integrals involving f(4) are combined. This yieids LEO} = Yn +O — fa — OI -— fO) + seo}. Remark. Since for the purposes of the Laplace transformation function f(F) is identified with /()S(), the term f(0) should be actually written as f(0 + 0). This reflects the fact that f(t) has, at ¢ = 0, a jump discontinuity as well (in general). 204 = THE LAPLACE TRANSFORMATION 58 The above argument can be trivially extended to a finite number of jump discontinuities of £(1). The generalized derivative theorem. If /(i) is piecewise continuous with, at most, a finite number of jump discontinuities in (0, 00), if f() is of ex- ponential order «, and if f°(s) is piecewise continuous for 0 < 1 < on, then LPO} = FO) ~ fO + 0) — y eG + 0) — fits ~ OF Example. Find £{/'(0} if 0 O@ 1). Since f(1) has a jump discontinuity at ¢ = 1, we use the generalized derivative theorem LPO} = sel} — (0 + 0 — eT + 0) — fl — Thus we have 2} = f eet) dp = where (i) is of exponential order unity. Also, f(0 + 0) = 9 fl + 0) = 1, and (1 — 0) = 0. Therefore (Res > 1, se? sal aif} = Remark, In this case it happens that ’(1) = f(0). 5.8 PERIODIC FUNCTIONS. RECTIFICATION Suppose that the Laplace transform of a periodic function f(t} (with the period T) is sought. It often happens that f(f) can be expressed by a simple analytic formula in the interval O < ¢ < 7, but not in the interval 0<1< o: For example, the saw-tooth function illustrated in Fig. 5.10, For such functions it is not difficult to evaluate the integral fay f 7 tp dt 0 as opposed to the direct evaluation of the Laplace integral =r -r T f ey) at. Figure 5.10 > aT 58 PERIODIC FUNCTIONS, RECTIFICATION — 205. To find £{f(0}. introduce a new function fe) OStsN, 0 @>T7). Let £{g()} = G(s). Suppose that g(/) is shifted by an amount F to the right along the faxis. The new transform is then G(s)e*? (by the shifting property). The transform of the sum, which represents “two waves” of the original f(#), is then Gi) + G@e*". Continue this process and obtain the transform F(s) of the periodic Function /(#) as the sum of the infinite series a= F) = G6) oe” a= The series can be summed (geometric series), mrt, l-er resulting in F(s) = G(s\/(h — e7*7). An alternative proof is to shift the function /(c)S(O to the right by an amount T and obtain the new transform F(s)e*7. Subtract this from F(s) which will result in F(s)(. — e*7). This is nothing other than [vod = 6), and the formula follows (this process repeats the method used in summing the geometric series). Certain periodic functions have their values in the second half of the period related to those in the first half of the period. For instance, sin ¢ has a certain behavior in® < ¢ < wand repeats it with an inverted sign in x < ¢ < 2x, namely, sin (+ x) = —sin#. Suppose that a function f(s) possesses the property fE4T/2) = —{0 (from which it also automatically follows that /(¢) is periodic with period 7). We define ow =f TH tpi) dt. o ‘Then it is readily seen that F(6) = G()/(1 + e777). Exercise. Perform the derivation of the above formula. Remark. Functions satisfying {(@¢ + T/2) = —JS(4) are sometimes called antiperiodic. 206 ©THE LAPLACE TRANSFORMATION 59 In many physical applications, antiperiodic functions are “rectified” by defining AO =f O<1< 7/2), * Salt + T/2) = Salt). This is usually calied the full-wave rectification and the new function is periodic with period 7/2. An example is the rectified sine wave in Fig. 5.11. fo Figure 5.11 It is not difficult to show that 2k fa(0) = coth T en, where G05) =f’ 7? O98) dt, 5.9 THE MELLIN INVERSION INTEGRAL The problem of finding £~'{F(s)} can be approached, in a general fashion, by the integral representation of the step function discussed in Section 5.9. It is known that £{S(¢ — 7)} = e77*/s, and tote = fe SD = a5 L eds. From this it is easy to construct the Laplace transform and an integral repre- sentation of a rectangular pulse, defined by ; O O 0. In order to evaluate the integral uf = Evi ote fO= Tri ten 5° eds, we close the contour (for ¢ > 0) as shown in Fig. 5.15, with the cut along (— 0, 0) because of the branch line of function /S [and, therefore, of function F(s)]- Em- ploying the technique of Section 2.15, Example 4, it is possible to show that the contributions from the arcs MP and P'N of the circle Cr vanish in the limit R > 0. On the segment PQ we have* s = re’ = ~r and v/s = rel? = ir, while on the segment P’'Q', s = rew* = —r and V5 = re? a —i/r, We find that the contributions from these segments are af Imi Sr (— R ine try uf 1 seur tre ae eat sf TAO CO R * _ if sink tg, 2 f sin test) ay ato r alo * In the limit e— 0, See Section 2.12, Example 7. 210 THE LAPLACE TRANSFORMATION 5.10 (setting k\/r = u and taking the limits p > 0, R— o). It remains for us to consider the integral over the “small circle” C,. Because of the polelike behavior of F(s) ats = O, we expect a contribution from C, that is analogous to a residue.* Indeed, using s - pe®, +e 1 eggs L __ i iG ds x |, I Consequently, by use of the residue theorem (no singularities of F(s) inside our contour), ahi 1 avs 2 fTsinu went a Lertetds = 1-2 J SN cutie ay, o Bri Syste 5 ado ou The integral can be evaluated without much trouble and reads 2 f * sin taney? k =] ——e du = erf—- rio wt Consequently, a1 fl | k £ ft = erfe —- 5° wt Exercise. Evaluate the above integral. [Hint: First evaluate Jo cos xu- e+" duand then integrate both sides with respect to x.] 5.10 APPLICATIONS OF LAPLACE TRANSFORMS Example 1. In the LRC-series circuit shown in Fig. 5.16, the switch S is closed at 4 = 0 and opened at ¢ = J. We want to find the current i(/) if (0) = 0 and (0) = 0. Kirchhoff’s law requires that B rE + RQ + D= en, HER Cc iO) i. where ' —ao) et) = len On. ° Moreover, from the definition of current, we Figure 5.16 find that . . ag ‘ 0-4. * It should be emphasized, however, that s = 0 is not a pole of F(s), but a beech point. (Compare with p. 96.) 5.10 APPLICATIONS OF LAPLACE TRANSFORMS 2u1 Apply the Laplace transform to both differential equations, i.e., multiply by e—* and integrate from 0 to oo: LsKs) — Ei} + (1/C)0G) = Ets), 46) = 5Q(s) — 90). Use initial conditions and eliminate Q(s): (Ls + R + 1/Cs)s) = E(s), where T toe Es) = eof edt = eg —*—.- A $ Therefore af tne 1) = TSE (R/Ds + VEC Leta = R/2L, and w? = 1/LC — R2/AL?, Three different cases atise: a) The oscillatory case, w® > 0. Inversion of I(s) yields itt) = & e-*"' sin wt — eum sin w(t ~ TSU — 1). b) The overdamped case, w? < 0. Let w? = —f7 and replace w by i8 in the above formula; i() = ne” sinh pf — no sinh A(t — T)S(t — T). ©) The critically damped case, 0” = 0. The transform reads ¢o_t-e*" 1-1 GERM so that i) = e te — 2 — Te SEE — T). Example 2.. A beam of rigidity ET is clamped at one end and is loaded as shown in Fig. 5.17; the weight of the beam is neglected. The deflection »(x) is sought. For the coordinate system shown, the following differential relations are known to hold.* 4 to hol Y d?y(x) 1 a) gee ey OD, wae where p(x) is the deflection of the beam at point x, and m(x) is the bending moment [counterclockwise Figure 5.17 v * For example, Salvadori and Schwartz, Differential Equations in Engineering Problems, Section 2,10, 212 THE LAPLACE TRANSFORMATION 5.10 torque of all (external) forces to the right of point x], de) b) ~y = tx), where 1(x) is the shearing force (resultant of all vertical forces to the right of point x), a) co) Ee = 9), where g(x) is the load per unit length at point x. From these relations it follows that (as on p. 174) wf) = 4x). Since the beam is clamped, we have p(0) = 0, y'(0) = 0. Also, 1(0) = gofZ/2), as well as = GLE _ _ gob? m= — "a TB From these last two conditions, it follows that Y"O) = ok? REF yO) = —9ol./2EI. ‘We now transform the DE and use all four initial conditions to obtain — Gel? , gol _ gg Le? 39) — 8 ger + 36F ~ Es so that ¥) = go eo? , gohk? 1 gol 1 # # EF ss} BE si 2&7 st Inversion yields _ fot of, 8) 5 -F gol? 2 Gob 2 YO) = yEt* — F4ET ( £)' s(x + the?” ~ Ge It is more convenient to rewrite this solution in the form hgh Lt go Lx Lx Ly, Be Tt ) (0 Sx< %) (le - a) G M?. The usual condition in practice is L\Z_ > M?. Thus we have e Las + R; 1 ~ Gy My a = 706 = 75) ny Me, WO = ~ Gy MH EE 214 9 THE LAPLACE TRANSFORMATION 5.10 where r, and rz are the roots of (kg ~ Ms? + (LiRa + LaRis + RiRe = 9. Since (LRe + LeRi)? — 4RyRoEiL, — M?) = (LiRe — Riz)? + 4R,R2M? > 0, both roots are real. Moreover, they are both negative since EyRo + LoRy Rie Tl, — M2 >o and Ll, — M2 > 9d. Rational Iraction decomposition yields 1s) = —£2 [#2 Lei +R: Lora + Re |. a Eyly — M? lryras "rir — ras — 71) * ralte — Ms — re) ——_Meo__.__1 a ty. RO) = —TE WP A-nbkon poh Inverting (using ry72 = RiRo/(LiL2 — M)], we find that {eae + Re yt _ Lata + Re et, TL, — M8 bry = r2) rary ~ F2) + Me jt 0O- Gia wD Tr © n= xt nto ety It is seen that 4; — €9/Ry and #; —> 0 as £— 9 as expected on physical grounds. Also, i,(0) = 0 and i,(0) = 9. The case of LiL, = M?, although an idealization, reveals an interesting point of theoretical importance. The determinant of the system is now a linear function of s, A= (E1Re + EaRis + RRs, and the transforms of the currents are @o/s(L2s + Ro) Meg 1O- G+ + BR POm ~ CR FR FRR yielding . 0 eokrRe RiRo iO = Ri, Rake + RS (- [Re + Ea ): (i) =< — fod __ _ _ RiRe . : WO = — T+ GR p( Tike + GR, ‘) i Note that these two functions do not satisfy the initial conditions i1(0) = .f2(0) = 0. Instead, we have i= colt éobz __ 4 . EyRz + L2Ry Ree GR 7) 5.10 APPLICATIONS OF LAPLACE TRANSIORMS: 215 Before we inquire into the causes of this discrepancy, we may convince ourselves that our results constitute the true solution of the physical problem. Jn fact, the conditions (0) = 0, (0) = 0 are inconsistent with our system of differential equations: Lt di; ait Mag + Rit = e05(), wii 4 1,22 4 Roly = 0. Indeed, let us multiply the first equation by M, the second one by L,, subtract, and use the condition LiL. = M7. The result is MRi(t) — LiRaia() = Meo — (for £ > 0), from which it follows that i(¢) and 2(¢) cannot simultaneously approach zero as > 0 (but ¢ > 0). Our physical intuition then fells us that at the moment the switch was closed, the currents i, and ig have instantaneously jumped to certain “‘nitial” values, from their previous values #, = fg = 0 {which are certainly valid for 4 < 0). To find what these values can be, we integrate either of the differential equations from —e to +e, where € is a small positive number. Then, since i;(—e) = 0 and io(—€) = 0, Mi,() + Laial) = 0. As ¢— 0, the values i,(€) and i,(€) approach the “adjusted” initial values of the currents: lim #,(€-) ~ 10 +0), Sim ie) — i2(0 + 0). Se 3b ‘Therefore the adjusted initial values must satisfy the simultaneous equations Mi,(O + 0) + L2i2(0 + 0) = 0, MRO + 0) — LyReio(O + 0) = Meg. Solving these, we obtain 40 + 0) = 73 _ tole eo UR + ER, nO+O= ~ TR TER, exactly as in our solution. This analysis helps to explain why we have obtained the correct solution with the incorrect initial conditions: The adjusted conditions are effectively buift in within the differential system describing the state of alfairs for ¢ > 0. This phenomenon of “instantaneous” adjustment of currents occurs quite often in switching problems in electrical networks if certain idcalizations are made (ike the assumption Z,Lz = M? which is, strictly speaking, unattainable in physi- cal systems}. Similar phenomena occur in mechanical systems, subject to sudden blows which change their state of motion. 216 THE LAPLACE TRANSFORMATION 5.10 Example 4. Consider the DE xy” + y’ + xy = 0 with the initial conditions yO) = 1, yO) = 0. In this problem, the formula apy) ~~. 6) - 2009) can be put to use. Using the initial conditions, obtain £7") = PIG) ~ 5, so that £{xy"} = —2e ¥(s) — FO 4 I. * S The transformed equation reads -2Y—s*¥4+14+s¥-1-¥'=0, or Yii+s*)+sY¥=0. The solution of this DE is elementary: YQ) = C//T FS, where C is a constant of integration. To find y(x), expand Y(s) in inverse powers of s: Cc 1\? _c | E(i+% =5|! ” oF ey 1-3-5-- Onn) | 4+ =ce cry 2m i Peat cat Dan(atp srt ¥() {Laurent series converging for |s| > 1). Assume that this expression can be in- verted term by term. Then (pte vay=c Lee Tingle ‘To justify the procedure, observe that the series for p(x) converges uniformly for 0 < x < Rwith an arbitrary R. Therefore, it may be multiplied by e~* and inte- grated term by term, yielding the serics for ¥(s). In other words, it Follows that ¥(s) is indeed the Laplace transform of p(x). Adjusting the constant C so that y(0) = 1, we obtain the ucsired solytion which is the Bessef function of order zero: Jolx) = Ec D" semaye x, PROBLEMS =. 217 Remark. This method may be helpful in solving some other differential equations of second order, but it may turn out to be useless if the transformed equation is more complicated than the original one, Also, if the solution has a singularity at x = 0, the initial values y(0) and y‘(0) cannot be specified (for instance, Yo(x), the second solution of the above equation, cannot be obtained by our method). BIBLIOGRAPHY Carstaw, H. S. and J. C. Jaecer, Operctional Methods in Applied Mathematics, New York: Oxford University Press, 1947. Cuurcuitt, R. V., Operational Mathematics. New York: McGraw-Hill Book Co., 1958. Scorn, E. J., Transform Calculus with an Introduction to Complex Variables. New York: Harper, 1955. VAN DER Pot, B. and H. Bremmer, Operational Calculus Based on the Two-sided Laplace Integral. New York: Cambridge University Press, 1955, PROBLEMS 1, Show that the functions cosh af and sinh at are of exponential order and evaluate their Laplace transforms, 2, Evaluate £{\/7 et}. For what values of s will the transform exist? 3. By performing the rational fraction decomposition, establish the results given below. afs+1 , ag Letts = 1+ sint — cost af osti |, 2 wm v9 by £ bratty -' var sin“5-# 4 od 2 aifst+2%42] 17 1 Pie 3 oe (gp2ee2l te tye -gta-G 4, Solve the following differential equations by subjecting them to a Laplace trans- formation, as in the flow diagram in Section $.4. ay'y ty = e%, — y(0) = 9 by" ay + dy = 0, yO) = 0, yO=t OW -v¥-w=P, yO) =1, YO=3 dy" + dy = 0, yO) = 1, yw/4) = -1 [Hiat: Treat y'(0) in (d) as an unknown parameter and determine it at the very end.] 5. Show that the following systems of differential equations have the solutions indicated: 2 oat ‘dx, dp _ ibe tar ata 7 = hs xO) = Of 20 = ——,—_—_ dy 2 1+2- Y_ye =O -5rtn xt at 3y > 0) vy 4 218 a x y THE LAPLACE TRANSFORMATION tBu ag x0) =1, 20) =0 » i ade " 7 N= \d’y dx . az t 2% -— ZVE-% vO)- 0, 30) ~0 x() = Jeosh 2¢ — 4 cos 2¢ v) = XP sinh 2 — ve in 21 {Hine: Apply the Laplace transform simultaneously to both differential equations and solve the resulting equations for the transforms x(s) = £{x(} and Y(s) = £9}. Consider the DE y” — 3y' + 2y = x? with (0) = ¥, »’() = §. Show that \Gee-¥), Observe what would happen if the expressions 1/(s — 2) and i/(s — 1) are ex- panded as in Section 5.1 and comment on the significance of the initial condition chosen above. |. Let Nat), Na(t}, and Ne(t) represent the number of nuclei of three radioactive substances which decay according to the scheme AM BBC, with decay constants X4 and Xp (the substance C is considered stable), Then the functions N4, Nz, and Ng are known to obey the system of differential equations Ng ay ae _ 7 AANA, do dae + daNa, ao” ABN. Assuming that W4(0) = No, and Wp(0) = Nc(0) = 0, solve the problem by Laplace transform methods and show that . Ae ket pa apt), Ne(O = w(t wou tien Also calculate Na(/) and No(s). Show how the convolution theorem can be used to establish the following result: i lo 5 (remer + Asinar) 2 af? 2 ke a . Solve the problem of a damped harmonic oscillator (as in Section 5.6): HAD + wbx = ((, x0) = x0, 10) = 00, i ‘by means of the convolution theorem for a) ws <7 (overdamped case) b) wi = 0? (critically damped case) 10. PROBLEMS, 219 Establish the following results: {CIC} = —(1/2s) log + 5, L{EI(-) = —(1/s log 1 +5), where cis — f Se du {the cosine integral), t ‘ EM = f Gt (<0) (the exponential integral). - Use the theorem on integration of transforms to establish the formuta 1 — cesar! I a of=sq] - zte(1 +9) . Starting from the formula © {erfe (k/2V7)} = (1/sle~*¥3, develop the inverse trans- form of e*¥*, [Hint: Use the differentiation rute.] . Using the result of Problem 12 and employing the rule of differentiation of transforms, develop the inverse transform of (1/0/s)e-*¥, Using the primitive function theorem and the result of Problem 13, develop the inverse transform of (1/s\/s)e~#¥7_ . To find &—'{F(s)}, the function F(s) is sometimes expanded in a series of suitable functions and inverted (erm by term. For instance, show that 1 foi I+k © * fake [ewe2e tt. = ete AD | sinh V/s % ° wt wt by expressing F(s) in terms of exponential functions, using the expansion ane and inverting term by term. This idea is widely used in heat-conduction problems. } Illustrate the Mellin inversion integrat by applying it to the following inversion problems: LMG? + a}, SO/Va. Specify the contour in each case and produce the necessary mathematical details to develop the results. Rg - In the circuit shown (Fig. 5.19), the switch S is closed at © = 0 when (he charge on the capacitor is g(0) = go = const. Set up the equation satisfied by q(i) for 1 > Oand 94] show that f a(t) = que—"foos wt + (@/a) sin of}, as s where a = R/2L and 0? = 1/LC — R2/4L?, Figure 5.19 220 THE LAPLACE TRANSFORMATION Figure 5.20 Figure 5.21 18. in the cireuit shown (Fig. 5.20), the switch S is kept closed for ¢ < 0 and is opened at #0, Assuming a steady DC-current prior to = 0 (eo is @ constant EMF), set up the equation satisfied by i(t) for ¢ > 0 and solve it by methods of the Laplace transform to show =a eo __ O-RERt RRER RAR VULEL NS 19. Consider the simply supported beam of Section 4.8, Example 2. In accordance with the statements on pp. 174 and 212, one must solve at: i) = a(x) = 9 = const, subject to »(0) = y(L) = 0 and y"(0) = y“(L) = 0. Observe that these conditions are more suited to the Fourier series method than to the Laplace transform method. Nevertheless, the problem is manageable if we use the Laplace transform [see the hint in Problem 4(d) above). Find y(x) by the Laplace transform method and show that it possesses the Fourier sine series displayed on p. 174. 20. The switch S in the circuit shown in Fig. 5.21 is kept open for ¢ < 0, the capacitor is kept uncharged and a DC-current (due to e9 — const) is present. If the switch is closed at ¢ = 0, show that the Joop currents 11(¢) and i2(t) and the charge q(t) on the capacitor obey the following relations: da dig -4. LG t+ Rat Ze, LE + Ra 2-0, where dy/dt'= i, ~ in, 110} = i2(0) = e0/2R. Solve this system by Laplace transform methods and show that e0c ~at oat at) = (1 e~"cos wt — De sinat) 60, 60 oat NO = Fat graf inet ; i £0 at oe bf) = # Foe sin wt, where a = R/2E and w? = 2/EC — a?. Caution: Before inverting, simplify the transforms Q(5), F145), and Jo(s) to the utmost degree. CHAPTER 6 CONCEPTS OF THE THEORY OF DISTRIBUTIONS 6.1 STRONGLY PEAKED FUNCTIONS AND THE DIRAC DELTA FUNCTION In physics, we often encounter the concept of a pulse of “infinitely short” duration. For instance, a body set in motion (from rest) by a sudden blow attains a momen- tum equal to the impulse of the blow, namely, mo = T= f tot Fa dt, where F(1) is the force and 7 is the duration of the action of the force. The term “blow” implies that 7 is so small that the change in momentum occurs instan- taneously. However, since such a change in momentum is a finite number, it follows that F(1) should have been infinite during the blow and zero otherwise. Force FU) Figure 6.1 This kind of description is not proper in terms of common mathematical con- cepts. For that matter, it may not even be physically rigorous. Indeed, the actual graph of force is more likely (o be a strongly peaked function, as in Fig. 6.1, where his very large while r is very small such that the area under the curve is equal toa Biven value of J. In many cases, a great majority, as a matter of fact, the exact shape of the strongly peaked function [F() in this case] is not known. However, insofar as the observable physical effects of such functions are concerned, this lack of information does net usually matter. What is significant, though, is the intensity of the impulse, namely, the value of the integral f tot" Fu) dt, p 221 222° CONCEPTS OF THE THEORY OF DISTRIBUTIONS 61 ae) Fx Jao dx idealized Actual Figure 6.2 and the time when the impulse occurred, namely fy {or, perhaps, fn + 7/2, but that hardly matters if 7 is sufficiently small). Strongly peaked functions are common to all branches of physics. For instance, a concentrated force acting on a beam is actually a strongly peaked distri- bution of load (Fig. 6.2). In electrical circuits, strongly peaked currents of ex- tremely short duration often occur in switching processes, like the redistribution of charges between the two capacitors shown in Fig. 6.3 when the switch S is closed. Initially, the voltages VW; = Q;/C, and V2 = Q2/Co are assumed to be different. When the switch is closed, there is a rush of current through it until the charges Q, and Q2 are redistributed into g, = AFG), gy _ CHO. + Os), 4 are 2 Or $e Wf the resistance of the leads is negligible, then this current pulse is of infinitely short duration and the current is infinitely large. Needless to say, this cannot be rigorously true; apart from the inevitable resistance (small, but never zero), there will also be a self-inductance L of the loop which will tend to moderate the steep rise of the current to its peak value after the switch is closed. In short, the current pulse will be a strongly peaked function of time. i L s QnM A Oh”, T 1" Cc Cy Figure 6.4 i In order to facilitate a variety of operations in mathematical physics, and particularly in quantum mechanics, Dirac proposed the introduction of the so- called delta function 5(x) which will be a representative of an infinitely shaiply 6.2 DELTA SEQUENCES = 223 peaked function given symbolically by _ {9 #0), 1" | =O) but such that the integral of 8(x) is normalized to unity: f * 50x) dx The first and basic operation to which Dirac sought to subject 8(x) is the integral { * seoysen dx, where f(x) is any continuous function, This integral can be “evaluated” by the following argument: Since &(x) is zero for x # 0, the limits of integration may be changed to —e and +e, where € is a small positive number, Moreover, since f(x) is continuous at x = 0, its values within the interval (—e, +e) will not differ much from (0) and we can claim, approximately, that [5 senses) de = [™ seoyftn dx = AO] Bd de, with the approximation improving as € approaches zero. However, + f ix) = 1 for all values of ¢, because 8(x) = 0 for x + 0, and 5(x) is normalized. It appears then that letting €— 0, we have exactly c Bo) dx = (0). Remark, The limits — co and + may be replaced by any two numbers aand & provided a In) n/2 x} < 1 /n)} form a delta sequence. Indeed, consider the integral [2 e090) ax ona) = @ = 1,2,3,..) for any continuous f(x). By definition of ¢,(x), we have [2 ease ax = (7 anges = oy" sas. - Now, using the mean value theorem for integrals, we can deduce that ofr ” fla) dee = (n/2)Q/a)f@) — FO) (- Vn S$ ES HA/n). If #— «, then §-> 0. From the continuity of f(x), it follows that f() > f), establishing the result Jim [" b(2)/la) dx = 0), which qualifies ¢.(x) as a delta sequence. For many, purposes it may be desirable to construct delta sequences from functions which are continuous and differentiable (this was not the case with the above sequence), For instance, some such sequences are n 1 a) n(x) = © T+ ee’ dG) = ; ©) a(x) = 2 Sita, : x 6.2 DELTA SEQUENCES = 225 All these functions are normalized 10 unity in the sense* that re n(x) dx = 1, and each sequence possesses the desired property [provided f(x) is continuous and the integral converges}: tim f°" bn fC) dx = £0). But again, it is incorrect to say that these sequences converge to the delta function: The limits of these sequences do not exist (according to common definitions of convergence). Remark. The normalization of the ¢, is not mandatory for a delta sequence; it should only be required that tin [Decor = as follows from the sifting property by setting f(x) = 1. Exercise. Show that limy_,0 ¢a(x) does not exist for any of the three sequences fisted above. tl) tad Figure 6.4 It is instructive to investigate the indefinite integrals of functions ¢,(x), namely, the functions a(x) =f dnll) de. For instance, for the sequence (a) given above, the graphs of ¢,(x) and ®,(x) for rather large values of n would look like those shown in Fig. 6.4, Also, it is not difficult to verify that lim 4,2) = S@), * There are two commonly used definitions of this term. For the second definition, see p. 468, The integral need not be over (—-20, ++<0) but rather over an interval appropriate for the problem under consideration. 226 CONCEPTS OF THE THEORY OF DISTRIBUTIONS 63 and we may be tempted to say that 5(2) is the derivative of the step function S(x). Again, however, dS(x)/dx = 0 for all x # 0 (which is all right), but the derivative does not exist at x = 0, exactly where the peak of the delta function is expected to be. 6.3 THE &CALCULUS The use of delta sequences made of differentiable functions has a very important consequence; for instance (Fig. 6.5), let dn(x) = (n/V/me—""*; then db) Pate deo Ve Consider the integral he $0 o [02 no where (x) is differentiable. Integrating by parts, we obtain * 42 pe [#2 pear = ecorooltt - [4.0L ae Assume that lim (a//a)e"™ f(x) = 0. Jim, afl aye fx) 65 This is usually true since we are deating with functions for which the integral S22 n(x) S(3) dx converges.) Then, letting 2 > 00, we have 40 im im fe dol) ayy dx = — fin J delalf'@) dx = - FO. It appears, that the sequence $4(x) is associated with the derivative-sifting property. This gives rise to the symbol 8/(x), the so-called derivative of the delta function, such that [iO 80opeyax = — 70). This idea can be developed further, giving rise to higher-order derivatives of; 5(x) possessing the property Cr ID 00) de = emo. . 63 THE 6-CALCULUS = 227 It is needless to emphasize thal the only meaning that can be assigned to the above statement is that Fo om m. firma [. FO) 10) ae = cy, and it is tacitly assumed that the functions involved are differentiable om times and that the integrals £ “Fonte fixax converge for all # and for all k from 0 to m. Jt should now be clear that the treatment of 8(x) and its derivatives as functions in the ordinary sense (which they cannot be) is a shortcut method for obtaining results depending on certain limiting processes. This procedure may be called é-calculus and it is widely employed in physical literature. As in the case of Heaviside’s operational calculus, it witl yield valid results provided its limitations are recognized and observed. We may develop the various rules of §-calculus by formal analytic operations starting from the properties a) "sa de = 0), 6) 20) - £ se, : -(° (x #0), D8= |. =o) with [Coan and ignoring, for the time being, their mathematical justification. Example 1. Verify the rule that x 8(x) = 0. Consider [12 x 3G0f()dx, where f(x) is continuous st x= 0. Write xf) = g(9); then g(0) = 0. Consequently, rr x 8(x)flx) dx = 0 for aif continuous (at x = 0) functions f(x). This justifies the equivalence of x 8(x) and the zero function. Example 2. Determine the meaning of 5(x — a). Consider [72 s(x — a) f(ddx. Set x—a then Eand write f Q and write, using ax = £, dx = (1/a) dt: d{ax)fix) dx = {- 8 E/a\(/a) dé = (1/a),f0). Ifa < 0, use ax = t,dx = (1/a) dt again; now, however, the limits of integration are interchanged and [0 Map) dx = i a()AUE/a)(1/a) dé = —(1/a) (0). In either case, the result is (1/Jal)f(0), thus establishing the rule. Remark: From this it follows that 8(x) is an even function (set a = —1). Example 4, Verify the rule 8G? — a?) = (1/20) + a) + (x —@)) (a> 0). Observe that 5(x? — a) = 3{(x + a)(x — a)}. Since 4(£) = 0 unless €=0, it follows that é(x? — a?) = 0 except at the points x = -La. Therefore, we can write [ yo? — a(x) dx = fo Ax + ax — af) dx +f Ge + ae — afiddx (a> 0), where 0 < € < 2a and € can be arbitrarily small, Now, in the neighborhood of x = —a, the factor (x — a) may be replaced by ~2a. Then .; ete i Ale + ay — a(x) dx = [ ag Maye + al fix) dx mate =f _. [al 8(x + a) f(x) dx te -{- 3g 8 + afta) de. ‘The infinite limits can be used again because (x + a) = Oexcept at ¥ = —a, 64 REPRESENTATIONS OF DELTA FUNCTIONS — 229 In a similar manner, ate + [ _, Hee + ax — aif) dx = [ eg 8% ~ MG) dex, and the rule is established. Remark, This rule breaks down for a = 0. There is apparently no way of interpreting the expression 8(x3). 6A REPRESENTATIONS OF DELTA FUNCTIONS For the purpose of formal operations employed in the 5-calculus it is convenient to establish for &(x) such expressions as Fourier series, the Laplace transform, and integral representations. Consider the pulse ¢,(x) shown in Fig. 6.6. Its Fourier coefficients dre b,=0 (the function is even), ay = £/E, a, = (1/nwa) sin (nxa/L) (rt > 1). Therefore the Fourier series reads 1 <1. nwa MRX tla) = ap + z pea ST COS =F v- Letting a — 0, we obtain the Fourier series for 8(x): x &) -L leas. +l +h Pot onx w= at 7 en: Figure 6.6 Observe that this series is divergent; this should not come as a surprise because if it were convergent, then 5(x) would exist as a bona fide function, which as we know, is not the case. Despite this drawback it should not be concluded that this series is absolutely useless. For instance, multiply it by a “well-behaved” function f() and integrate term by term from x = —Ltox = +L: 40 4h t St mx, 21S xz}, Soydxst Le |, Pecos “Ede = + 2 a where a, (n = 0,1, 2,...) are Fourier coefficients of f(x). If (x) is represented by a Fourier series fi) = +E (cncos BE + sin 4). then, i€ (x) is continuous at x = 0, (0) = B+ Y an. aa 230 CONCEPTS OF THE THEORY OF DISTRIBUTIONS 64 It Follows that the divergent series ate LL, cos — does possess the sifting property and may be treated as a representative of the delta function. In a similar fashion we may develop various other series for delta function. For instance, the Fourier sine series for 5(x — £) reads seo? 3 sin MEE sin BE @ 0), , o-t 65 APPLICATIONS OF THE &CALCULUS 233 where e may be taken to be arbitrarily small. Such a force f(2) can be approxi- mated by the function Fat — to) so that the differential equation to be solved reads E+ IE + wor = (I/m) 8 — to) (A = p/2m, « = Vk/m). Suppose that the oscillator is at rest at ¢ = 0; we tacitly assume that fp > 0, Then 0) = 0, 40) =0. We shall solve this equation by means of Laplace transforms, The transformed equation reads 2X8) + DsX(5) + 3X5) = U/mje** so that entto XO = ETF aE The Toots of the denominator arer= A + 4/n? — ah. To be specific, assume that wg > d? and set w? = «2 — 7. Furthermore, write 1 nGoAeomy™ XG) = and invert by use of the Heaviside theorem: fo KE—tod +o") x) = if Hom tear Now user) = — + i, rp ~ —A — iw and reduce this to X(t) = U/me™"— sin wt — to)St — to) or 0 (t < te), U/mje "© sin w(t — t9) (E> to). x(t} = It is instructive to obtain this solution by conventional methods. Since the force f(t) supposedly acts instantaneously at ¢ = ig, the oscillator obeys the homogeneous equation for both ¢ < tg and ¢ > fy: ¥+DE+ whe = (t % ft). Also x) = wu) <0 (for # < to). For > fo, the general solution may be written as x(t) = Ae~™ sin (we + ¢)- 234 CONCEPTS OF THE THEORY OF DISTRIBUTIONS 6.5 To find A and 4, note that physical conditions imply that x(1) is continuous everywhere (including ¢ = f9). Also observe that this cannot be said about x(/). With this in mind, integrate the DE from tg — € to fo + €: foe eae eaafeh att of edt = Im tone toe tome or Rio + €) — Xo? — €o) + WxCto + 6) — xl — + 02 f Oh" dt = Him. tome This is valid for an arbitrary «. Let ¢-+ Q and use the continuity of x(1); then lim f'** (de = 0, and — lim [xp +) — x(t — = 0. : wo 0 Hye Note also that X(f9 — €) = 0 (both x and % vanish for f < fo), Consequently, lim X(t9 + €) = X(to + 0) = I/m. ed Since 3(0) = —de™MA sin (wt + 4) + we™ cos (wi + 4) (E> to), we can obtain, letting 1+ f9 from the right, T/m = we™A cos (ote + 4). On the other hand, letting f+ fo in the expression for x(t), we can obtain 0 = eA sin (wig + 9). Thus, sin (wtp + $} = 0, oF wip + @ = nx (n = integer). Then cos (lo +) = (—® and A = (—1)"(E/omme*”. ‘The solution reads x(t) = (—I"Ufame™—" sin fot — tp) tne} (E> to) Whatever m may be, the above result is identical with “ x(t) = Ufeamye*""0) sin w(t — tg) (> fo) It is seen that the solution obtained through é-cafculus is a correct one, and that we arrived at it with fess effort than by using conventional methods. ‘As another example, consider a simply supported beam loaded by a con- centrated force P at the center (Fig. 6.10). In the DE* i ! dy) _ it Pee FI), * See p. 174. 65 APPLICATIONS OF THE 5-CALCULUS: 235 we can now set g(x) = P (x — L/2) on the P grounds that P actually represents an infinitely sharply peaked distribution of load with [eM ode P, > 0, tin where ¢ is arbitrarily smafl. This problem can be solved efficiently by the Fourier sine series expansions. Assume that v2) = 3 by sin (nex/, and use the (divergent) series ein MX 2 FR yy tt gi BEX a(x- 1) a? PEt sin“ = 5 z ( 1y sin? n=a.3, Substitute this into the DE, Mk 2P St cn—ty ft gj BRE nx St 1 sin "32. EG a TE BIL... L Consequently, mowig PL? Lg by = } Ehet na = 000), 0 (1 = even} and QPL’ Rex 90) = Fie sin It is not difficult to verify (by methods similar to those used in the first example) that this is indeed the Fourier sine series of the correct solution y(x). Observe, however, that a direct attempt to verify this will fail, we will hit a snag after dif- ferentiating the series three times, at which time uniform convergence will cease to be valid. Remark. As in Example 2 of Section 4.8, we find that the series is rapidly convergent. For instance, if x = L/2, the leading term alone yields the approximate deflection mw 2 PE G = which is only about 1% off the exact result 236 = CONCEPTS OF THE THEORY OF DISTRIBUTIONS 66 6.6 WEAK CONVERGENCE The ideas mentioned in the preceding scctions which lead to the concept of the delta function can be systemized into what is known as the theory of distributions, or generalized functions, As the name suggests, the theory is concerned with the problem of extending the definition of a function so that concepts such as (x) can be put on a firm mathematical basis. There are various ways to achieve this, We shall pursue the idea expressed in Section 6.2, namely, the approach to distributions through integrals of sequences of functions of the type SAGO) de (= 1,2,3,...) A sequence of functions f(x) (such as a delta sequence) leads to a new mathe- inatical concept (such as delta function), provided such a sequence of integrals converges for any suitable function g(x). Now, what shall we call “suitable?” From Section 9.3 we know that if we wish to define concepts like 8'(x), 8"), etc, then g(x) should be infinitely differentiable.* Also, we shall assume that the.limits of integration (unless stated otherwise) will be —co and +o. Therefore, g(x) must have “proper behavior” at infinity. We shall insist on a very strong require- ment of this type, namely, that g(x) must be identically zero outside some finite interval (a, 5) (a different one for different g(x), in general).t Functions g(x) that satisfy these requirements will be called test functions. This name is appropriate, since the sifting property of delta sequences, for instance, is tested on these functions. Example, The function HH Ut < a) 0 dal ©? G a= { is a test function. Indeed, it is evidently infinitely differentiable for |x} # a. As for the points x = -ta, one can show, using the basic definitions of the derivative, that all derivatives exist and are equal to zero, As an illustration, let us demon- strate that g(x) = 0 at x = —a. For convenience, we seta -+ x = £ and express gasa function of £. In the immediate vicinity of the point £ = 0 (that is, x = —a) we have 9 {ES 0), a) = {rma (>0), By definition of the derivative, dg(x) = 42 _ jig RD = 8 _ tim &@). Cr ort) to £F * That is, differentiable any number of times. ~ t Less restrictive conditions can be adopted, requiring only that g(x) approaches zero “sufficiently fast” asx — 4:00 (¢.g., Lighthill, Fourier Analysis and Generalized Functions). 66 WEAK CONVERGENCE 237 Wf & <0, then g(¢) = 0 and the limit on the right-hand side of the equation is zero. IE > 0, we use 7 = 1/£ and write lim (1/807 222-® = firm ge! 2a— A, ra) soe Since e72 (2a) <=! and since lim,_,, ¢—2" = 0, it follows that the proof is complete. Having defined the test functions, we can now proceed to define the class of core functions* from which the functions f,(2) will be selected. There is a variety of choices} Unless stated otherwise, we shall require the core functions to be infinitely differentiable over the entire range (— 2, +0). Their behavior at infinity may be arbitrary. ‘The term core functions is related to the fact that it is this class of functions which is being extended to encompass other (not infinitely differentiable) functions as well as bona fide distributions, such as the delta function. Now consider a sequence of core functions f(x) ( = 1, 2, 3,...). It will be called a weakly convergent sequence provided that the Jimit im f™" fulade(a) dx exists for all tes? functions g(x). A weakly convergent sequence may or may not be convergent in any of the conventional definitions, like pointwise convergent, uniformly convergent, convergent in the mean, etc. The concept of weak convergence is a special one designed to extend the class of core functions in a peculiar fashion. Such extensions are possible by means of other types of convergence, but, in a sense, extension by weak convergence is more powerful. Consider, for instance, the sequence of core functions fal) = 4+ (I/x) arctan nx. It is not difficult to verify that this sequence converges (pointwise) to the step function S(x). However, the core consists of functions infinitely differentiable over (-~, +0) and S(0 is not a core function. Therefore, the sequence f,(x) cannot be said to converge “within the core.” Nevertheless, it converges weakly: One can show that sim f° fated a = [™ scapgtey dx = f° a) dx for all test functions g(x). We are now in a position to define artificially the discontinuous function S(x) by adjoining to the core functions the “limit” of f,(x) defined only inside an integral * There is no uniform nomenclature in the theory of distributions. Most authors will call these functions adauissible functions. + There is no unique theory of distributions, but rather there are several, based on similar principles but differing in details. 238 CONCEPTS OF THE THEORY OF DISTRIBUTIONS 6.6 by means of lim f * 4 + (/ryaretan mxyg(x) dx = f * Scog(x) de. Why do we try to define the step function S(x) in such a complicated way? Is it not much simpler to define it pointwise as follows?* 0 <0), 3) = ( & = 0, 1 (@>9). The answer is obvious because, with exactly the same ease, we can define the delta function which defies any pointwise definition. Consider, for instance, the sequence of core functions quoted on p. 224: n 1 SO Ea As n> co, this sequence does not converge to any function (within the core or otherwise). Nevertheless, it converges weakly and dim [face dx = 80) for any test function g(x). The proof is as follows: to n t UVR n 1 i a DE EO ee “f TF Was 80) rey won Yoav eT mB AOE + Yd Tp maT ed Let 2 be the bound for g(x), namely, [g(x)| < B for all x. Then n I t 1 < af 2 Teme t= a(;- 4 arctan Vin) f at ax) dx par Te Since lim (@ —Laretan va) =o 208 it follows that tim f 2 de = 0. nove Jive DH ax 82) de = 0. * Thereby extending the core by pointwise convergence. 6.6 WEAK CONVERGENCE 239 Similarly, vA ' - a lim f 7 Tbe a(xjdx = 0, Finally, using the mean value theorem (f.(2) > 0). +a The 1 i xipe a(x) dx = «of “wel Thm” = a(t)? arctan Vn CUVn SS 41/0), so that ELEN 1 . n Tie faye TE wixe BO) & = 80), completing the proof. We can now give a rigorous definition of a distribution as follows: A distribu- tion ¢(x) is a mathematical concept associated with a weakly convergent sequence of core functions* for which the symbolic integral [& serge de has a meaning, by means of the prescriptiont [> eet dx = fim f™ fone as. Remarks 1. Each core function corresponds to some distribution since we can construct a se- quence consisting of only a single core function f(x). 2. A great variety of different weakly convergent sequences will yield the same values of +. lim f Slog) dx for all test functions g(x) (e.g., various kinds of delta sequences). Such equivalent sequences correspond to the same distribution. * There is another, more abstract, but also more general, approach lo the theory of dis- tributions which defines a distribution as a linear continuous functional over the space of test functions (e.g., Gelfand and Shilov, Generalized Functions, Academic Press, 1965). This definition dispenses with core functions altogether, along with the integrals. + For convenience in describing the relationship between a distribution (x) and the weakly convergent sequence {f,(x)}, we shall use the expression “the sequence (f,lx)} weakly converges to (a distribution) ¢(x).” The precise meaning of this statement is given by our definition. 240° CONCEPTS OF THE THEORY OF DISTRIBUTIONS 67 Example 1. The following sequences are equivalent, all corresponding to the distribution denoted by ¢{x): a) fQ) = Ferfe(—nx) = awn” edu, by) fi) = 4+ (1/e) Si) = (/n) f~ (sin u/u) du, ) f=, Note: Sequences (a) and (b) also converge pointwise to the conventional step function S(x) (as given on p. 114); sequence (c) converges pointwise to the function 0 <9, SQ) =| l/e (x= 0), 1 (> 0). From the point of view of classical theory of functions, S(x) and 3(x) are different Sunetions. From the point of view of the theory of distributions, S(x) and 3) correspond to the same disiribution (x) which is not even defined by a formula like that for S(x) and 3x), but rather through the limiting process deseribed in weak convergence. Example 2. The sequences a fA=2 Kea atest e _ int ©) fala) = RESO, gy) pgp Sine, are equivalent, yielding the result rte tim f™ fede) dx = 2(0) (all g(@)). noe tow The distribution defined by any of these sequences is called defta function* and is denoted by a(x). 6.7 CORRESPONDENCE OF FUNCTIONS AND DISTRIBUTIONS Let ${2) be a distribution. In this case the expression [2 enewax ; is not an integral (in the Riemann sense), but rather the limit of a sequence of * Of course, we should call it delta distribution instead, but the name delta function is now part of a fong tradition, 67 CORRESPONDENCE OF FUNCTIONS AND DISTRIBUTIONS — 241 Riemann integrals. However, it may be equivalent to a Riemann integra! if a function {(x) can be found (not necessarily a core function) such that [2 Perec) de =f 66a)e(2) dx = lim f"* KGa) ae + for alltest functions g(x). Asa rule, this will happen if the sequence f,,(x) converges pointwise to some function f(x) which can be taken as f(x).* However, sur~ prisingly enough, there are exceptions to this “obvious” rule. Example. The functions e-MN—Aine—071 y — U/n| < 1/22), F(X) = { 0 (x = i/n| 2 1/20) are evidently core functions.f They are infinitely differentiable everywhere and yanish outside the interval (1/2n, 3/2n). Let us normalize them: FO) FO) = Fae rua (s0 that now 72" (0) dx = 1). The sequence of functions f,(x) converges pointwise to zero. Indeed, if x < 0, then f,(x) = 0 (for all n). If x > 0, choose N large enough so that W > 3/2x. Then f,(x) = 0 for alla > N and the statement follows. Observe, however, that the convergence is not uniform. The sequence f(x) also converges weakly. For any test function g(x) we have, by the mean value theorem (note that fa(x) > 9), + 2128 [7 Sabereta) dx = fo fale) dx = BOO" falerde = gO) 12a SES YP), Now, if »— oo, then £ must approach zero (from the right). By continuity of test functions, , lim g() = g@)-t boo * Clearly, f(x) is not unique, as the case with Sx) and S(x) shows (Section 6.6). We have {- Sle) de = J" Sede ax for all test functions g(x). + These functions are of the same kind as g(x) om p. 236, with a = 1/2n, and centered about the point x = 1/n (rather than x = 0); that is, x has been replaced by (x — 1/n). { This means that (/,(0} is a delta sequence. However, while previous delta sequences were divergent, this one is convergent (to zero); convergence has been achieved by shifting the peak of f,(x) while making it higher and narrower. 242 CONCEPTS OF THE THEORY OF DISTRIBUTIONS 67 We have, therefore, shown that lim fe) = F(x) = 0 so that CC fim fe) dx = 0 [all eC), but din {** f@)s(a) dx ~ g(0) [all gO) This example illustrates that the concept of pointwise convergence may have very litle meaning from the point of view of a physicist. Indeed, if functions Sn(%) represent successive approximations to a certain physical quantity and the formula re Salt) dx = 1 Temains true at all times, how is it possible to reconcile physical intuition with the statement that f(x) “converges to zero?” It is reasonable to conjecture that, at least for certain purposes, distributions are more suitable for description of physical quantities than conventional functions. Let us look into some problems created by this proposal. First of all, it should be firmly remembered that we cannot talk about the value of a distribution at any given point.* For instance, 6(x) cannot be assigned any value at x = 0. By itself, this would not be so bad, but the fact that the entire contribution to the integral °° iQ) dx = 1 comes exactly from x = O is rather embarrassing. As another example, the dis- tribution 6(x} [corresponding to the step function S(x)] is defined by the sequences of the integrals Jim f° S.Gde09) de = [fF eas, and does not distinguish between S(x) and S(x) of Section 6.6. It is equivalent to either of them. Evidently, #(x) cannot be assigned a definite value at x = 0. Does not this property of distributions contradict the commonly assumed properties of physical quantities as functions? Perhaps not. For instance, con- sider the measurement of an electric current (2) varying in time. The current is measured by an ammeter possessing a certain inertia: It cannot respond to in- * From that point of view, the notation $(x) for a distribution is misleading. The symbo! (x) is meant to indicate that there is a number y [y = ¢(x)] corresponding to the number x. We follow, however, this traditional notation which has ils obvious advantages, 67 CORRESPONDENCE OF FUNCTIONS AND DISTRIBUTIONS. 243 stantaneous changes in current. What is being measured is not really a function A(t) but rather some sort of average current, more like the expression ef 1a) dr for some finite « (however small it may be). From this point of view physical quantities hardly possess the mathematical properties of classical functions. Nevertheless, since classical functions certainly approximate physical quanti- ties, it is necessary Lo show that distributions can perform the same task. fn other words, we must show that there exist distributions virtually indistinguishable from ordinary functions, at least of the types used in physics. The major step in this direction is accomplished by the folowing theorem. The smudging theorem. For every continuous function f(x) one can con- struct a sequence {f,(x)} of core functions so that | fal) — FOI] <€ for arbitrarily small ¢, uniformly in x within any finite interval. Proof. Consider the sequence of the core functions* je Um MO—BY eg] < tf) Yates = | 0 x — > U9. They will be normalized if we divide them by the normalization integral aut ; stm Mn =f cxo|- Te wil -fhn, op (3 me) & (by a change of variables f = x — £); note that Nj is independent of x. Now define a sequence Tein piay= ff” rotate, 94, lin which is a sequence of core functions. Indeed, f,(x) is infinitely differentiable itself and vanishes at x = +1/#, along with all its deri es. + See the footnote on p. 241. + The formula af i a ) dats) P bx w Jaq FONE = f wo a Ut Sle bOI -Z — fle, eG] is being invoked. It is valid in a finite interval [e(x), b(x)] if 3//8. is continuous. 244 CONCEPTS OF THE THEORY OF DISTRIBUTIONS 67 Remark. What is being done is the averaging (or smudging*) of the function f(x) over the interval (x — 1/n, x + 1/n) by means of the function U,.. This creates an infinitely differentiable function f(x) out of f(x) (which may lack such a property) in such a way that fn(x) is a very good approximation to f(z) for large values of #. We can also write Sp lin eA fn z—tin Un(x, £) a t Foye Wala 8 at = OUACIE zn FX) = f) and construct the expression (since U, > 0) [fa) — f0)| = 1/0 hts | sin LE) — PON UaGs, 9) ab. Since [x — 4 < 1/n, one can always choose n large enough to make [x — ¢| arbitrarily small which implies, by continuity of (2), 4@ — FQ) < € valid for all x in an arbitrary finite interval, with € being arbitrarily small. Then ehtin UG) — fool < Mf" UG, Dae = € Ain as required. The uniform convergence provision in the smudging theorem permits us to formulate the following corollary. Corollary. For every continuous function f(x) an equivaleny distribution (x) can be found such that [Eee nato ae = [" Ket ae for any test function g(x). Indeed, let f,(x) be a sequence of core functions as described in the smudging theorem. Let g(x) be a test function; by definition, it must vanish outside a certain interval (a, 5). Therefore [[2 fade ae — J" sede de| = | {Unto — sooletw as| [A - folleeol ae By the smudging theorem, |f,(4) — f()| can be made arbitrarily small forall x inside the interval (a, b). Therefore ? [Stents de ~f" parecer de| < ef lacalax = Bb — «, * See Lighthill, Fourier Analysis and Generalized Functions, p. 22. 68 PROPERTIES OF DISTRIBUTIONS —- 245, where Bis the bound on [g(x)|. Since we can make the right-hand side arbitrarily small by choosing # sufficiently large for any given test function g(x), it follows that . te0 tim f"* facoe(o dx =f fone) de, establishing the equivalence between a distribution g(x) defined by the left-hand side of the above formuia and a continuous function f(x). Remark. The equivalence of f(x) and (x) can be extended to piecewise continuous functions. However, it is simpler to treat these functions as derivatives of piecewise smooth functions* and employ the differentiability property of distributions (see the discussion in Section 6.8). 6.8 PROPERTIES OF DISTRIBUTIONS Distributions can be subjected to various finear operations common to ordinary functions. First of alt, distributions can be added and muttiplied by constants {in other words, linear combinations of distributions are also distributions). This follows from the consideration of sequences of integrals [2603 + AGO dx and [TtchGoee) dx (C= consy. Tf f(x) and #,(x) are core functions, so are the functions Sn(X) = fa(X) + fn) and 2x) = Char). Moreover, sequences s,(x) and z,(x) must be weakly convergent if the sequences Jn{x) and 4, (2) are weakly convergent. Therefore, the above sequences of integrals define the distributions o(x) = 6 + XQ), FX) = COO), where {x} and x(x) are distributions corresponding to f(x) and A,(x). Distributions can also be multiplied by infinitely differentiable functions. In- deed, let A(x) be an infinitely differentiable function and Jet the sequence f,(x) be weakly convergent. Consider the sequence [7 weap otets) ae. * See p. 167. 246 CONCEPTS OF THE THEORY OF DISTRIBUTIONS 68 The functions ya(x) = A(x)f.(x) are core functions, Also, the above integral must converge as 1 —» 00, because it may be written as [O seoiec2 ax, and B(x) = A(x)g(x) clearly is a test function. We can therefore define a new distribution (x) = A(x)6(x) corresponding to the sequence yu(x). Remark. It is not possible to define a product of two distributions in a general fashion (e,, it is not possible to define a product of two arbitrary distributions*). ‘The reason is that if a sequence of functions pa(x) = fa(x)in(2) is formed, it may not be weakly con- vergent and does not define a distribution. For instance, lett Sule) = tl) = C/V aE? It is not difficult to find a test function g(x) for which the integrat te [2 fob ee) ae will fail to converge {e.g., @(x) = 1 for |x| < @ and arbitrary otherwise}. Distributions also permit linear transformation of an independent variable, namely, if #(x) is a distribution, then @(x — a) and (Cx) are also distributions (where a and C are constants). They are defined by the sequences of core functions f(x — a) and f,(Cx) which are weakly convergent because one can write [00 fe — angterax — [* facet + a) ae and [°° ecaeen de = elf Kerex/ode — (C ¥ 0). Since g(x + a) and g(x/C) are test functions and f,(x) is weakly convergent, the statement follows. One of the most important properties of distributions is their infinite differentia bility, The derivative of a distribution ¢(x) is associated with a sequence of deriva- tives of core functions fa(x), which converges weakly to ¢(a). To verify that this is possible, note that f(x) is a core function if f,(x)is one. Furthermore, integrating * Some particular distributions can form a product; ¢.g., the above example ¥ = fb can be treated as such a product since 4(x) is equivalent toa distribution, - } We are attempiing here to construct [6()]?; this turns out to be impossible within the framework of the theory. na asa eect ats tn Eg He 68 PROPERTIES OF DISTRIBUTIONS — 247 by parts, we obtain {c SiCOg(x) dx LGC — f" fate’ dx =f Falone'( ax, u because g(x) vanishes outside some finite interval. Since g(x) is a test function (if g(x) is one), the limit (as 11 —> 0) of the right-hand side exists. Then the limit of the left-hand side must exist as well, Therefore (x) is weakly convergent to some distribution which we define to be the derivative of ¢(x) and denote by ¢’(x). The distributions (x) and ¢’(x) are related by the formula [2 #0ogc ax = — {one ax for any test function g(x). If (2) is equivalent to some differentiable function f(x) (see p, 244), then ’(x) can be shown to be equivalent to f’(2).* Suppose, however, that $’(x) is not dif- ferentiable everywhere. To be specific, assume that f(x) is continuous and piece- wise smooth. Let us investigate in what sense ¢’(x) extends the definition of the derivative f’(x) to those points where it does not exist in the classical sense. First of ail, ¢’(x) does not really define ’(x) at any point: A distribution has no value ala point. What is still true for a distsibution is that it can be equivalent to a function in the neighborhood of some point, say x = t. This is shown as follows: Let (2) be a test function, different from zero only within some neighborhood of point £, namely, the interval (¢ — €, + 6. Now, if for a function f(x) and a distribution o(x) it is true that [PP econ ae = [poor ax Ene (aay for all functions B(x) of the described type, then we say that #(2) is equivalent to {00 in the neighborhood of point & If f(x) is continuous in (£ — , £ + ), then (x) can be constructed, e.g., by the method uscd on p. 244 {corollary to the smudging theorem). We can now show that if (x) and f(x) are equivalent at x = Ef and f’(z) exists, then ¢’(x) and f'(x) are equivalent at x = £ Indeed, if B(x) vanishes outside (¢ — «, £ + ©), so does #’(x), and we have, for alf such 3(3), pe (EO) dx -{" #'COR(X) dx = =f s0r'00 te the - =~ foe) dx * By a technique similar to that used in proving the smudging theorem. + For brevity, if ¢(x) is equivalent to f(x) in a neighborhood of £, then we shall simply say that $(x) is equivalent to f(x) at & 248 CONCEPTS OF THE THEORY OF DISTRIBUTIONS 68 However, exactly the same relation applies to f(x), yielding [pcan dx = — [Spoor ae. Therefore [econ dx = [°° Peano ax, establishing the equivalence of 4/(x) and f'(x) at x = £. ‘What about those points where /’(£) does not exist? Since f(x) is piecewise smooth, f(x) must suffer a jump discontinuity at x ~ £, and be expressible (in a sufficiently small neighborhood of £) by FRO =AOD+ SE-B O&O, where f(x) is continuous and Ais a constant. Since 9’(x) must “faithfully” follow S'(@) for x ~ &, it must satisfy the relation Bee [Eo senna ax =f" 09 + Sta ~ BRO dx, for all appropriate B(x). This relation uniquely defines ¢’(x) and makes it equiva- lent to the discontinuous function FQ) + AS(x — B. Note that it makes no difference whatever how S{x — £) is defined at x = € (see p. 240) because this fact does not affect the value of the integral [Wie + a8 - Oe ae. From this analysis we see that dist tions can represent discontinuous functions. {at least those with jump discontinuities).* Since the derivative of a distribution is well defined without any restrictions,t it follows that a “distribution with jump discontinuities,” as ¢'(x) above, can be differentiated as well. Let y(x) be equivalent, in the neighborhood of the point £, to a discontinuous function of the type PU) = ux) + ASG — BD, where u(x} is differentiable in (¢ — €, E+ © except, perhaps, at the point x = itself. A typical such function is t shown in Fig. 6.11. f Figure 6.11 * Many other types of discontinuities can be accommodated as well; see, e. Bs Gelfand and Shilov, Generalized Functions, Vol. 1. t From this it follows, of course, that distributions are infinitely ditferentiable. 68 PROPERTIES OF DISTRIBUTIONS 249. Let @(x — £) be the distribution corresponding to S(x -- ¢). Then +0 bo [2 #0 — eeydx = — fox — Dero ae iF = ~[ ste ~ pe ae ~[F ¢@) ae = - 200 = a0. From Section 6.6 it follows that "(x — £) corresponds to a delta sequence, and therefore (x — £) = B(x — 2). It easily follows that the distribution ¥'(@) must have the form* VO) = 1'G) + hx — BD, where u’(x), in turn, is of the general form wee) = u(x) + kS(x — Ek = const), 41(x) being some function continuous in (¢ — «, E+ oe. This analysis establishes the rule: When a distribution with jump discontinuities is differentiated, each discontinuity at x = ¢ of magnitude & (height of the jump) results in a term 8(x — £) in the expression for the derivative. Since distributions can be differentiated, they can be required to satisfy ditfer- ential equations, Because of the finear character of the theory of distributions, the differential equations satisfied by them should, as a tule, be linear, namely, of the forni ty = f, where x is a finear differential operator with infinitely dif- ferentiable coefficients.t Suppose that fis a bona fide function. Then the above DE can be viewed asa “classical” DE (for functions) or as a DE for distributions, Without going into details, we shall illustrate some of the features of the second point of view. If the DE has no finite singular points,t then it can be shown that it has no other solutions {in the space of distributions) except “classical” ones, ie., those with a counterpart in the space of functions. In particular, the DE dy . dx has, in the space of distributions, the general solution yp) = C = const. The presence of singularities in the DE can create, however, some peculiar situations. * The rule [ap(x) + x(x)" = a¢'(x) + 5x’) is taken for granted. The proof is triviat. t See p. 245 on multiplication of distributions by functions. 1 For second-order differential equations, see Chapter 3. For other cases, consult any advanced textbook on differential equations. 250 CONCEPTS OF THE THEORY OF DISTRIBUTIONS 69 Example 1, Consider the DE ay * ae for x > 0, p(x) must coincide with a constant. The same is true for x < 0. How- ever, the constant need not be the same. Indeed, the “step-function” distribution y(a) = 9(x) satisfies the DE because 6’(x) = 5(x) and* x a(x) = 0. So far as the distributions are concerned, our DE has two linearly independent solutions, say yi) = Vand y2(x) — 0(x), and the general solution reads y(x) = C1 + C202). Note that the classical counterpart of 6(x), namely S(x), cannot possibly satisfy the DE because it has no derivative at x ~ 0. Such a deficiency cannot happen to a distribution. Example 2. Consider the DE x? (dy/dx) = —y. It has a classical solution y(x) = 0 which is also a distribution. It also has a classical solution yQ) = Cel (C #0). However, this solution does not qualify as a distribution: No sequence of core functions can be found to be equivalent to Ce'!* for x + 0 and to be weakly con- vergent.t - Finally, differential equations satisfied by distributions may not be applicable to ordinary functions. The most commen and most important of these are of the type fy = a(x — 8, giving rise to the so-called Green’s functions (see Chapter 12). 69 SEQUENCES AND SERIES OF DISTRIBUTIONS In many applications we have to deal with sequences of distributions $(x) (m = 1,2,...) of, more generally, with distributions $,(x) depending on an arbi- trary parameter u. A sequence (¢,,(x)} of distributions is said to be convergent if there is a distribution (x) such that} sim {- dned@) dx =f” (2)9(0) dx, or, more generally, Tim f* dsCee(a) dx = f°" o6)e) be tng F—0 f + aed de = 0 Gail). ; { This is not because the integral [42 el/"g(a) dx is divergent (some divergent integrals possess valid interpretation in the theory of distributions); it is rather because it is badly divergent. ft may be shown that if Hirt, +012 daldeCo dx exists, then ox) must exist as well. 69 SEQUENCES AND SERIES OF DISTRIBUTIONS 251. for all test functions g(x). We write then, by definition, 40) — fim 4,09 0 and call @(x) the fimis of the sequence {443}. Example. The functions {cos mx} may also be looked on as distributions. In this sense they form a convergent sequence as m—+ oo. Indeed, fet g(x) be an arbi- trary test function, Then it must vanish outside some interval (a, 6) and be bounded inside this interval: leo SB @Sx< 6). Then fon 6 { cos mx: g(x) de = f cos mx: g(x) de = a & FT in nx - g(x) dx. Now, g(a) = g(6) = 0, while [ff sin mx: 2a) dx] < Bt — a), As we let m —> co, the right-hand side of our equation approaches zero. Therefore, the sequence of distributions {cos mx} converges to zero.* Note that the sequence of functions {cos mx} does not converge to anything: it is a divergent sequence. We have here, evidently, a different definition of convergence: For distributions, this concept is an extension of weak convergence of functions (as opposed to point- wise convergence), For that matter, the sequence of functions {cos mx} con- verges weakly (to zero, of course). Theorem. If (x) = lim, .294)(2), then $'(x) = lim ¢.(x). rae Proof. Recall that me : [o senedr = ~ [song's de. ‘Therefore His f° aeCaretaydx = ~ tier [* a00d9'(@) de ponte do posta dan JP aeneteads = f° eeneco ar. and the theorem is established. * By zero we mean the (unique) distribution called zero distribution, equivalent to the function f(x) = 0 in the neighborhood of any point x. 252 = CONCEPTS OF THE THEORY OF DISTRIBUTIONS 69 This property of distributions greatly facilitates various analytic operations with them. It does not hold, in general, for ordinary functions. For instance, the sequence of functions f(x} = sin mx/m converges to the zero function a8 m — o, namely, lim f(x) = fQ), — where fix) =0. me Now, cach function f,(x) is differentiable, but the sequence of derivatives, fida) = cos mx, is a divergent sequence. Note, however, that (x) possesses a derivative, namely, the zero function again. The convergence of a series of distributions can be defined, in a customary fashion, by the convergence of a sequence of the partial sums. For example, given aset of distributions ¥(x), k = 1,2,3,..., we form a sequence of partial sums: * G(X) = z WO) (m= 12,3...) If this sequence converges to some distribution o(x), we say that = ¥ vaG. k=1 Such a series can be differentiated term by term. Indeed, from lim om(x) = o(x), it follows that lim o3,(2) = a(x). However, ox) = D1 Hx). a Therefore oa) = fim Fo VAG) = De vi) as stated. Example. We wish to demonstrate that the formal series 777 is divergent in the usual sense, can be regarded as a distribution. ... cos kx, which Consider the series Cfo1a5,.. (L/k?) cos kx. ‘This series converges uniformly in the conventional sense and represents the periodic function ; 3/8 + ax/4 (—w Sx <0), GQ) = § 47/8 — 4x/4 OX x <0), {+ nx) otherwise (1 = integer). 6.9 SEQUENCES AND SERIES OF DISTRIBUTIONS 253 We can associate with /(x) a distribution (x) equivalent to f(x) everywhere (in a neighborhood of any point x) which may tr be considered as the sum of the above series within the space of distributions.* A(x) = x (fl?) cos kx. 13.5. Figure 6.12(a)_ represents graphically either g(x) or f(x). The series for $(x) can be differen- tiated term by term and yields 6@=- S Ofsinks, rat'3%5,.. ow representing the distribution so ‘7 +n/4 (-1 2E because the region where U(x, £) is different from zero does not overlap with the range of integration (—L, L). For —2L $ x < 2 there is a partial overlapping} as shown in Fig. 6.13; from this figure it is clear that u(x) can be calculated for 0 < x < 22 by the formula wea) = UN)” PNG @ £2 snresenes0 dedy = f° scot dxf sne200) do = g1(0)g2(0). Note, however, that 6() is axially symmetric (invariant under rotations). The partial derivatives of a distribution p(x, y) can be defined in a similar fashion as for one dimension. The following formula holds: te pte $0 pte f . [ 20 2) g(x, pdx dy = — f . f. ale, yD dx dy. * There is no difficulty in interpreting a product of distributions in this case because the distributions refer to independent variables. 258 CONCEPTS OF THE THEORY OF DISTRIBUTIONS Of particular interest in practice is the result of the application of the Laplace. differential operator r, # ea = VW ant a8 toa distribution (x, y). For instance, it is possible to show that 80) = —(1/2e)V* log (t/r) or = VR Fy, where log (1/r) is regarded as a distribution. BIBLIOGRAPHY Erovety1, A., Operational Calculus and Generalized Functions. New York: Holt, Rinehart, and Winston, 1962. GELFAND, I. M., and G. E. SHILov, Generalized Functions, Vol. 1, New York: Academic Press Inc., 1964, Lica, M. J., Introduction to Fourier Analysis and Generalized Functions, New York: Cambridge University Press, 1960. ZeMANtaNn, A. H., Distribution Theory and Transform Analysis. New York: McGraw-Hill Book Co., 1965. PROBLEMS 1. Let (2) be an arbitrary function analytic in the upper half-plane. Show that tin f™* pute fa) de — J00), where 1 a $00) = 2 pa aba In other words, ¢a(x) is a delta sequence for such functions f(x), {Hint: Set 1/n = a and evaluate the integral by residue calculus; then let a —» 0.) 2. Show that fim {” dal8) dé = St) for the functions n 1 monte? é De Tipae Dee Se i . Using the sequence of rectangular pulses (as in Section 6.4); develop the Fourier series for 8(x — £) in the interval (~L, +2). » » a PROBLEMS == 259 . If the step function S(x) is developed in a Fourier series in (-L, +2), the result is Su) = 542 dein ttt. weott E Differentiating term by term, we can obtain the series (2/L) Dtmeu Cos (n1nx/L). However, this differs from the Fourier series for 8(x) given on p. 229. Can you trace the reason for this discrepancy? Show that the step function admits an integral representation 1 fo e so- 3 f ot and specify the treatment of the path of integration at the pole. Give an argument in support of the formula 5(x)/x = —8'(x). What kind of rigarous statement can you produce in relation to this formula? . Consider the problem of the bending of the beam shown in Fig. 6.15. With the help of the discussion in Section 5.10 (Example 2) and Section 6.5, formulate the problem using the delta function. In particular, show that YO) = PEER -y'""(O) = —P/ET. Solve the problem by a method of your own choice and show that the deflection is given by Px? N we wa (E- 5) OS <8, pe vy) = : BE (. - §) Sx. Figare 6.15 . Consider the circuit problem of Section 4.8, Example | (Fig. 4.7). Suppose that the periodic electromotive force E(t) has the form of the Square wave shown in Fig. 6.16. Show that the problem may be formulated by means of the della functions, Solve the problem for the steady-state current i(#). Eo) [| le | ~T-34 -Th 4TA 3A OT Figure 6.16 CHAPTER 7 FOURIER TRANSFORMS 7.1 REPRESENTATIONS OF A FUNCTION A function f(x) may be represented, e.g., by a power series Se) =F ax" nad provided it satisfies the necessary conditions for this expansion. If the sct of co- efficients {a,} is given, then the function f(x) can be corstructed, at least within the radius of convergence. In other words, a function can be defined by means of the coefficients a,. Similarly, if a function f(x) is representable by a Fourier series ey Aay= Yee (-L oo is difficult to perform directly since c, apparently approaches zero. We follow the idea of using the frequencies. We set nx/L = k; then Ak = (m/L)An. The “‘adjacent” values of k are obtained by putting dn = 1, which corresponds to (L/r) Ak = 1. Then we can multiply each term of the Fourier series by (L/x) Ak with impunity and write #7, 10) = (Eos) ak, where [ +h Z .= x , Spe dy, Switching completely to the k-notation and writing (E/x)cn = ¢,(k), we can obtain 40 +” ek) Lf pear and ope FE extinet* ak. aed, tafeon The nomenclature employed here is borrowed from optics and is self-explanatory. 262 FOURIER TRANSFORMS 72 If we now let L —» «@ we expect the second sum to “go over” into an integral and we obtain 4° Gk) = Fim ex(h) = x E Aiaye™ dx and IQ) = fc cle dk. This set of formulas {in a somewhat different form) is known as the Fourier trans- Sormation. We shall adopt a modern yersion,* defining a function Fk) = Viz (—H, so that the formulas now read +0 ete fhe Fk) = Vel. Sexe dx, 1 ke =e Fie" ™ de f vin L ‘The function F(X) is known as the Fourier transform of the function SQ). The concept of the Fourier transform of a function f(x) is analogous to the concept of its Laplace transform. Conversely, f(x) is called the inverse Fourier transform of F(k) so that the transformation and inversion formulas are obtained at the same time. Note that the only difference between direct and inverse Fourier transforms is the sign in the exponential function. 7.2 EXAMPLES OF FOURIER TRANSFORMATIONS Jn the preceding section we have conjectured the validity of the formulas 1 f° Fk) = Velo Sexye™ dx, L{* Fe ak WR Strictly speaking, there is no question about the validity of the first formula. If £(3) is an integrable function and if the integral converges, F(k) is simply defingd by the above integral. The question, however, is whether the original function f(x) * There are several ways of defining Fourier transformation; the ‘differences between them are trivial. ° 72 EXAMPLES OF FOURIER TRANSFORMATIONS 263 can be retrieved by the second formula. We shall illustrate the validity of the second formula by considering some particular examples, Example 1. Consider the Gaussian probability function* S(@) = Ne (N,a@ = const), Ms Fourier transform F(k) will be denoted by ${f(x)} and will be calculated from +e +2 FO) = sf) = Te | foe de = aa [ oe dr, The technique used for the calculation of this integral was given in Section 6.4. We complete the square, ax? + ikx = —(a/a — ik/2/a)? — k?/he, and make the change of variables x\/a — ik/2\/a = u to obtain Nia ev due N Fa Vina —* 2a It is of interest to observe that F(k) is also a Gaussian probability function with a peak at the origin, monotone decreasing as k —» 4:00. Note, however, that if f(x) is sharply peaked (large a), then F(k) is flattened, and vice versa (Fig. 7.3). Thisis a general feature in the theory of Fourier transforms. In quantum-mechanical applications it is related to the Heisenberg uncertainty principle. F(k) = fey Fu) Large @ t x ae ee Fi t fo Small @ ee ef Ne Figure 7.3 * See Section 6.4; the constant NV is often chosen to be the normalization constant, yielding. S22 fo) dx = 1, 264 FOURIER TRANSFORMS 72 The inverse integral Lf” : tow of mean [eer x San Vie V2al—° tan be calculated by the same technique; as a short-cut, one may set a= 1/4a and x = —x; then 1 [" in’ he i 2, 2 ent Fett k ok thal Tq e- Vin im Via “e so that N 1 poe VE f. Fe dle =~ VPae™ = fix), and the conjecture is verified. Example 2. Consider the function f(x) = a/(x? + a) (@ > 0). Its Fourier transform is given by +> tke a e Fk) = $Uf(x)} = Valo pa This integral is just “tailored” for the calculus of residucs.* Note that the contour has to be closed differently for & > 0 and for kK < 0. Let us write e' 5 to ihe Fw -f Va @tan ay” If k > 0, the contour is closed “up”; the pole z = af yields the residue a ao Vie 2ai Fk)=vafie™ (b> 0). Ifk < 0, the contour is closed “down”; the polez = —ai yields the residue od a Vx (-2ai) (there is an extra minus sign because of the clockwise integration) so that : Pk) = Vale (k <0). so that * A similar problem was solved on p. 90. V2 EXAMPLES OF FOURIER TRANSFORMATIONS 265 For k = O it is possible to close the contour either way. Alternatively, using #4 etan®) > x? at dx \uctane we directly obtain +o FO) = 0 a 1 x = f. pa dx = = arctan 7 ‘The final resu't can be written as F(k) = Vrf2e** (all ky and is shown in Fig. 7.4. FUo Figure 7.4 Again, if f(x) bas a fairly sharp peak (small a), then F(i) is rather spread and conversely. Of some interest is the discontinuous derivative of F(k), which is related to the fact that f(x) does not approach zero sufficiently fast as x > 4-00 fe.g,, the Fourier transform for the function x/(x) is not defined for k = 0 and the Fourier transform of x? f(x} is not defined for any k]. Example 3. Consider now the box function ! (al a Its Fourier transform reads (a > 0), : 1 of 1 , Fk) = VE of dx = ee x to x The functions f(x) and F(k) are shown in Fig. 7.5, fe) Fk) =a | 4a \ aa Figure 7.5 266 FOURIER TRANSFORMS 73 7.3: PROPERTIES OF FOURIER TRANSFORMS Let us investigate the properties of the function F(A) = F{f(x)} defined by the integral t f ike kh) = A dx. FR) Vig J» Sexy dx. First of all, it is clear that F(k) is defined only if f(x) satisfies certain restrictions. For instance, f(x) should be integrable in any finite region. In physical practice, this means that f(x) has, at worst, jump discontinuities or mild infinite discon- tinuities.t Also, the integra! should converge at infinity. As a general rule, this would require that f(x) — 0 as x —> seco. A very common sufficient (but not necessary) condition is the requirement that f(x) is absolutely integrable, meaning that the integral [Oey as exists. Since [f(xe"*#] = L/(x)|, it follows that the integral for F(&) is absolutely convergent; therefore it is convergent. Because of the presence of the complex factor e**, it is evident that F(A) is, in general, a complex function.{ Let us assume, on the other hand, that f(x) is real, Then we have the so-called conjugation property F(-) = F*®). The proof is trivial: take the complex conjugate of the integral defining F(A). There are two immediate corollaries to this property.§ a) If (>) is even, F(X) is real, : b) If fo is odd, F(K) is purely imaginary. Exercise. Establish these two properties. ‘The appearance of the exponential function e** (sometimes called the kernel of transform) suggests that there should be similarities between Fourier transforms and Laplace trarisforms. Indeed, it is not difficult to establish the attenuation property and the shifting property for Fourier transforms: S{flae} = Fk — al), HI(x- a} = e* FH). } For example, the function {(x) = |x|""e~=" is integrable for 0 0 as x— oo. Then we have += fiye™| — 1 Pao - E fe dx. FLED) = HGF}. The extension to higher derivatives is obvious; in particular, the one most often used in physics is the relation TS") = - PFO. It is evident that Fourier transforms should possess many properties analogous to those of Fourier series.{ For instance, the conditions of existence of the Fourier transform are analogous to the conditions of existence of the Fourier coefficients. The problem of inversion of the Fourier transform is analogous to the problem of convergence of the Fourier series to the original function f(x). We have seen (Section 4.7) that this latter problem can be dealt with in the sense of convergence in the mean. If this is true for the inverse Fourier transform, then we should expect an analog of Parseval’s formula (p, [71). A bit of reflection (e.g., by a heuristic argument like the one in Section 7.1) tells us that we should expect that the formula [2 ear ae ~ f" ipcal? as, known as the (first) Parseval theorem in Fourier transform theory, will hold. ‘As a matter of fact, one may propose a more general formula: [0 FonG— 0 ak = [* fone ae, known as the second Parseval theorem in the Fourier transform theory. * Compare this with similar properties of the Laplace transform in Section 5.3. {Ff this does not hold, then f(x) can be treated as a distribution (see Section 7.5) and the differentiation property is still valid (p. 277). } For example, the conjugation property for Fourier series is given on p. 166. 268 = FOURIER TRANSFORMS 73 To establish this result, we use the definition 1 3 G-K = — [ —* des (-k) Vin Jan axe then +0 $0 + . Fk Hae =f Fekya f >= tt dy, |, FRG(-*) Ae) a aR Suppose we can interchange the order of integration, Then i FUQG(—k) dhe = [ . g(x) dx Suppose further that the Fourier inversion integral is valid: i [ * = Fie ck = f(x). x J, TOE Fe). In this case, the second Parseval theorem follows. The first Parseval theorem is then a mere consequence: Set f(x) = g(x) (real functions) and recall that G(-k) = G*( to obtain [© cago ax = [Ciscoe or 00, $00 [2 logan ae = f™ lg? ax, as stated. Remarks 1. A safficient condition for the interchange of the order of integration in the above proof is the absolute convergence of the integrals [rie Mae and [Oo coe ™ as, ; of, in other words, the convergence of poe 0 f \FG)}dk = and [ tea] dx. 2, The pointwise validity of the Fourier inversion integral is not reaily necessary: For instance, the values of the integral 17? = (ke ake aL Me 74 FOURIER INTEGRAL THEOREM 269 may differ from the values of the original funciion f(x) at some isolated points} provided this discrepancy does not affect the value of the integral [P ecoseyee. Exactly the same assumptions are sufficient to establish the convolution theorem for Fourier transforms. Theorem. If 5{4(x)) = FUQG(K), then HO) = Fae) = ef) = W/via) [** fee — Dat, where FGQ = {f(x} and GU) = {a0}. Indeed (denote the inverse Fourier transform by #7"), 1 fe S~{FGQGIQ) = VE f. FUNG Ke" dk 1 +? ik; +° ike = if” Fike af glee dt. If the interchanging of the order of integrations is permissible, then 1 _! + + —ike—b FFG} = 5 a(t) dé Fike dk 2x Jw 0 fo { = — dt, Viz [ FDS — Hat as stated, 74 FOURIER INTEGRAL THEOREM For those functions f(x) for which the inverse transformation is valid, we can write the integra! identity, known as the Fourier integral theorem: ie te Sa) = 5 L eW* dk Fbe™ dt. If f(x) is a real function it may be convenient to recast this identity into real form. We write ° - wk ae f ats f= Fe [rive ae + Fel, Reet a t This will actually happen (in general) if /(x) is piecewise smooth. 270 FOURIER TRANSFORMS 74 Making a change of variable, k’ = —k, in the first integral (and dropping the prime) and using the conjugation properly F(—k) = F*(k), we obtain 1 f uae + PY eyetts (x) = —— FRET Fk dk, Se) Vale FF@e (ke) However, 1 Vin Now, we take the complex conjugate of this formula to obtain F* (ke** and add: ary Fike"? = SDH ge. Fe" 4. Feit = 1 = [ SUB) cos k(E — x) de, Vie yielding the real form of the Fourier integral theorem: 0 ne SQ) = i f dk re Slé) cos k(t — x) dé. One of the basic mathematical Problems in the theory of Fourier transforms is to establish the class of functions £() for which this theorem is valid. The ques- tion is analogous to that in the theory of Fourier series: What are the functions Fe) for which the Fourier series will actually converge to f(x)? It is hardly surprising that the answer is similar: For instance, the following {wo theorems provide sufficient conditions for the validity of the Fourier integral theorem.f Theorem 1, If f{x) is absolutely integrable in (=o, +00) and is piecewise very smooth in every finite interval, then the Fourier integral theorem is valid in the sense that uff if af KE) 60s KE x) db = f(x +0) + fx — Oy). Theorem 2. The result of Theorem | remains valid if the statement "... is piecewise very smooth in every finite interval . . ."* is reptaced by “*. . . satisfies Dirichlet conditions in every finite interval...” We shalt omit the proofs of these theorems. As mentioned before (Section 6.7), the concept of pointwise convergence (involved in the above theorems) is not only unnecessary in physics but may be even meaningless. The idea of con- vergence in the mean (Section 4.7) or, perhaps, that of weak convergence (Sec- tion 6.6) is certainly more suitable. } Compare with the theorems in Section 4.6. Again, the above conditions are by no. means necessary: For instance, functions with certain “mild” infinite discontinuities also conform to the Fourier integral theorem. TS FOURIER TRANSFORMS OF DISTRIBUTIONS = 271 7.5 FOURIER TRANSFORMS OF DISTRIBUTIONS Among functions possessing Fourier transforms {and satisfying the Fourier integra! theorem) we can find those used in the construction of the delta sequences of Sections 6.2 and 6.6. For instance, the sequence of functions Inf) = iV ae generates the sequence of Fourier transforms” Fh} = UpV2ne Vi", We know that the sequence {f,(2)} converges weakly to (x). It would be natural to define the Fourier transform of the deita function by the limit Latent Jim eT" Vigne Of course, one should prescribe what kind of convergence is implied in the above limiting process. In this particular case it makes no difference because e—?!4"? con- verges to unity, both pointwise and weakly: lim Fa(k) = im e747? = | and Fim f°" e-®"*9(k) dk = f ** ok) dk nave Jom = for all test functions g(k). From this point of view we are inclined to claim that $(8(x)) = 1/2. Naturally, we are also interested in the possibility of inverting the Fourier trans- forms of distributions. This means that we will require 1! 1 fa uf? ax) = 1h ee dk = if oe dk. @ (a) Vie Ja Viz a). 4 This formuta has already been obtained on p, 230.} [ts obvious interpretation is that the sequence of functions +0 1 I —Rtyan?, ike wi. ° € dk converges weakly to the delta function as 1 — oo. This is of course true, since these functions obviously are the fq(x) defined in the beginning. + See Example 1 in Section 7.2. { The sign of the exponent makes no difference. Make a change of variable k’ = —k and drop the primes. 272 FOURIER TRANSFORMS 15 Hf we want to provide this idea with a general basis, we must make sure that, given a weakly convergent sequence {fC}, the sequence of corresponding Fourier transforms {F,(4)} is also weakly convergent. The immediate snag in this requirement is that many core functions F(x), as defined in Section 6.6, will not possess Foyrier transforms at all. For instance, f(x) = x2 is a core function, but the integral f + se dy diverges. If we have grasped the fundamental idea of the theory of distributions, we see that the way out of this difficulty is to restrict the core in sucha fashion that the core functions are sufficiently “well behaved.” This will cause no difficulties because such a restriction is not really a restriction at all. Just as we were able 0 generate discontinuous functions out of infinitely differentiable core functions (see p. 237), we should also be able to generate functions like SG) = x* out of our restricted core. The only difference is that such functions will now become distributions, but this is hardly a disadvantage (see p. 242). The core can be restricted in various ways. Perhaps the most convenient restriction for our purposes is as follows: The core functions f(z) should be infinitely differentiable over (— 00, +0) and, along with all their derivatives, should vanish at infinity faster than any power of x. In other words, tina [x"td"f(x)/dx"]] = 0 for all m and all 2.* . The advantage of this particular restriction is that not only the Fourier trans- form of each core function exists, but that it is also a core Sanction. To establish this, observe that +s ee = * ft ixfQde™ dx, The differentiation under the integral sign is valid because the integral [2 te dx converges according to the properties of f(x),t thereby establishing an absolute and uniform convergence of dF/dk. These arguments can be repeated any number of times, yielding 40 TF) _ 1 Fy ypike ; GO. 2 fede dx. ‘ * These are the “good functions” of Lighthill. T Note that x/(x) is a core function; for instance, Ixf@O} < B/x? ase + to: 76 FOURIER SINE AND COSINE TRANSFORMS 273 Therefore F(k) is infinitely differentiable. Now, integrate by parts and obtain PFE) _ O™ FO dmg y ite Te SOF Sel. ae Aixyle"™ dx. ‘The term which was integrated out is evidently zero. Again, this operation can. be repeated any number of times, say n+ I times, to yield om _ to peel er) = (ern Om f . oy sone a. Since x” f(x) must be a core function, its (n + 1)-derivative must be one too, and the integral on the right converges. From this, it follows immediately that ndF{k)t A dkm k for some constant A. Therefore 0 (allm,2), completing the proof. With this definition of core functions, the base is laid for the development of a theory of distributions incorporating Fourier transforms. It is not difficult to establish the Fourier and Parseval theorems for core functions and to transplant these (along with other) properties of core functions onto distributions generated by weakly convergent sequences.* The distributions generated by this core include many functions that are nonvanishing at infinity, such as polynomials of all orders. Although more re- stricted than distributions generated by the core of Chapter 6, they are adequate for physical purposes. 7.6 FOURIER SINE AND COSINE TRANSFORMS Let the function f(x) be an even real function. Using the real form of the Fourier integral theorem, we can write 1 fs) = ‘ f de} f2foos kx cos KE + sin kx sin ke] de te md 0 = 1 cos kx af AQ) sin ke ae pb f sinkxdk{ — fig)sin ke dt. eso «© lo — * Lighthill, Fourier Analysis and Generalized Functions. Lighthill uses test functions of the same type as core functions, calling them “‘good functions.” 274 FOURIER TRANSFORMS 76 Since f(£) is even, the second integral vanishes and the first one can be modified to yield fix) = 2f cos kx ak f Sle) cos KE de. rio o This. formula suggests the following idea, analogous to that in Section 4.4 for Fourier series: Suppose that a function f(x) is defined only in the interval (0, an). Jf we extend its definition over the entire interval (— oo, oo) in an even manner defining SQ) («> 0), nyo PO = on (x <0), then the above formula should be applicable to f(a). In this spirit, we define the Fourier cosine transform by* Fo(k) = Self} = AE LF 10) cones as with the inverse FEF} = Re f Fe{k) cos kx dx = f(x). Remark. In these formulas there is no mention of the function {H(x). However, if we choose x to be negative in the inversion integral (in the first integral, x is a dummy variable) we automatically obtain an even extension of f(x), namely, f(x). OF course, if we are interested exclusively in the region 0 < x < «, this remark is irrelevant.} Needless to say, the Fourier cosine transform should have its odd counterpart, the Fourier sine transform. If f(x) is odd, then the Fourier integral theorem yields feo = 2 f sink ef AO) sin ke de. eso 0 Consequently, if {(x) is given only over (0, 0), we can define Folk) = S500) = af? f flo) sin kx dx with the inverse 53 '{Fs(k)} ~ Viz f ” Fa(k) sin kx dk = fQ). : Evidently, the inversion integral reproduces f(x) in the “odd fashion” for (— 00, 0). * Since only the region x > 0 is involved, we can revert to the symbol f(a) instead of sPG). : t See, however, the convolution theorems on pp. 275-276. 76 FOURIER SINE AND COSINE TRANSFORMS: 275 Fourier sine and cosine transforms are expected to Possess properties similar to those of Fourier transforms. For instance, integrating by parts, we have Fel f' OY = V2/m [" (x) cos dx dx = V2/x f(x) cos kal + Viek f ” s(x) sin kx dx. In most cases of practical interest, f(00) = 0.* Then Soff) = ~V2/x fO) + kFs(k). Note the “interference” of the Fourier sine and cosine transforms. However, Jet us calculate 5¢{f'"(x)}, integrating by parts twice (assuming f’(00) = f(co) = 0); the result reads FoF") = —V2/x (0) ~ KF o(k) and involves only Fourier cosine transforms, Similar results are obtained for Fourier sine transforms. Our experience with Laplace transforms and Fourier series should tell us that Fourier transforms, Fourier sine transforms, and Fourier cosine transforms can be employed in much the same fashion, for instance, in solving differentiaf equations. The above result then indicates that the Fourier cosine and sine transforms are convenient under certain special conditions, such as in the absence of the derivatives of odd or even order in the differential equations and when relevant functions and their derivatives vanish at infinity.+ The interrelation between Fourier sine and cosine transforms is also exhibited in the convolution theorems. For the cosine transform we have BaF elk)GK)) = V2/x [" Fe(Gclk) cos kx dk = Q/x) f ” Fe(k) cos kx dk f ” g(é) cos ké dé. 0 0 We now use cos kx cos kt = #[cos k(x — ) + cosk(x + 8] to obtain (interchanging the order of integration) Fa'Fo(k IGRI} ~ (/n)f” a(6) at f” Fe(k) feos kx — 9 + cosk(x + Dldk ~ WIVE)L” OMG ~ B+ Set Ode. Note: The appearance of f“ (even extension of f) is due to the fact that in the Fangex S & < 0, wehave(x ~ £) < Oand the inversion of Fe(k) must yield f°. + See also p. 277 for a related comment in the theory of distributions. 1 The last requirement can be relaxed if distributions are used (see Section 7.5), 276 = «— FOURTER TRANSFORMS 17 If we perform the same operations for the Fourier sine transform, then we will use sin kxsin kt = f{cosk(x — §) — cos k(x + 8] and deduce that 5a '{Fs(RIGSAD} = (1/a){" aCe) def” F(k) [eos k(x — §) — cos ke + Rak. Note that the sine transform Fs(k) is now “mismatched” with cosine functions. ‘We can define a new function* LQ) = Fe" {Fa(K)} and state that 55 {Fo(K)G th} = W/V Inf” OU — D ~ Jee + Diab. Observe that /(x) is, by definition, an even function. - 7.7 APPLICATIONS OF FOURIER TRANSFORMS, THE PRINCIPLE OF CAUSALITY Example 1, Consider a damped harmonic oscillator acted on by an external force g(t). The motion of the oscillator is then governed by the differential equation $1) + 2ax() + wdx() = S(O, “where (0) = (L/m)g(). This problem has been treated in Section 2.5 for the case where f(s) is a sinusoidally varying function of frequency w. By means of Fourier transforms, we can extend this result to an arbitrary function JQ). In all cases of practical interest, f(t) will possess a Fourier transform; then Si) = US Inf ™ Faye" da, where Fa) = (\/V7e) f= fide" at, The solution x(t) is also expected, on physical grounds, to possess a Fourier transform [which we shall denote by 4(«)] so that x) = Vin) f * Aye do. ' * We transform f(x} by a sine transform and invert by a cosine transform, J At least as a distribution; for instance, any bounded piecewise continuous f(A) can be treated as a distribution according to the results of Section 7.5. 17 APPLICATIONS OF FOURIER TRANSFORMS 27 We can easily find A(w) by subjecting the differential equation to a Fourier trans- formation and using ${%} = —iw5{x}, and 5{%} = —wF{x}. Note: These formulas assume that x{--00) = %(t00) = 0. In many cases this will be true. However, if x(t) is treated as a distribution, no such restriction is needed. We have, by definition,* te +0 . 4 xe! di = ~ [ _ £ edt, The transformed differential equation reads —w? A(a) — LawiA(e) + wRA(w) = Fw) and yields Fw) Ae) = ——F@)_ = Eat) eal The solution of the problem is then 1 fren *O= Vix J—= (2 — w) — awl da. In most cases this integral can be evaluated by the calculus of residues. For this purpose it is necessary to know the zeros of the denominator,f Zw) = («2 — w?) — 2awi, because they will create the poles of the integrand. Ifa > 0, asis usually the case, the poles are in the lower half-plane and are located at wy = & Vuk — 0? — af if oo > @ (Case t), O12 =(-at Ve — adi fog ). The Fourier transform of f(t) reads te Fw) = z f edt = fo dx? then be ; iet _ fh / singe WO- ~ 2 |, ae ole = on™ where a, = p — di, 2 = —p — ai (p = Vad — a”). The only singularities of the integrand are those at @ = w, and w = ws. In order to find how the contour can be closed we must investigate the bounded- ness of the integrand. We write 0) sin wr = (1/2i)(e#" — e— #7), : + ; It is clear that if ¢ > 7, the function fo sin wt - e co 7 o k— r—afe—r is bounded in the lower half-plane and the Figure 1.7 WW APPLICATIONS OF FOURIER TRANSFORMS — 279 Im Im Path Figare 7.8 Figure 7.9 contour may be closed.““down” as shown in Fig. 7.8. This yields fo [eter ‘near th = + 2,;/b ere _ _ Simagr-e Om eer = wa) axles — 0) 2foi p ~iwyt (wa SiN wre onaaalea; — ws) & * = @, Sin wyr - eM8), Ifa real form is desired, set = p— oi og =—p— al (wwe = —w?) and express sin 7 and sin war in (erms of exponential functions: XO) = (Sofa felTHOMT=O 4. glemipnr—ty tet ipl 4 taint $+ (ai/ pyle Hmr-9 tam inem—b_ totione to 4 ee This expression reduces to (0) = (fo/abXeos ptt — 7) + (a/p) sin p(t — r)Je-!-? ~ Ulo/esoeos p(t + 1) + (a/p)sin p(t + rile"? (> 7), For ¢ < —7, the contour may be closed “up” and yields zero. This is expected: Before the external force has begun to act, a damped* oscillator could have been only in the state of rest. If | < 7, matters are somewhat more complicated. It is evidently convenient to split the integral for x(1} into two parts, according to the decomposition sin wr = (1/2i)e*" — e 7), The part containing e~*“'+” may be evaluated by a contour closed “down” and the part containing e—"'—” by a contour closed “up.” However, each inte- grand will then have an extra pole at a = 0. Since the “combined integrand” appearing in the original expression for 3()) has no pole at w = 0, it is clear that the path of integration may be deformed into the one shown in Fig. 7.9, without affecting the (overall) result in any way.t * That is, “positively” damped (case a > 0). 1 One can circumvent the pole from “above” just as well, or take the principal value. This will change nothing. 280 FOURIER TRANSFORMS qd tm iim First te nd irst term, Flewre 7.10 Second term After this is done, the combined integrand can be split as stated, with both parts assumed to be evaluated along the path given in Fig. 7.9: __ fe [ en ietn fo f eueen 10 = ~ 97 Jpste = wile = ay) ™ + Pei Jpowe = oie = aq) Now the contours can be closed as planned (Fig. 7.10), resulting in fo evieeey en feat) j= —-22 - ,4__. ~_. *O- ~ os Poe a5 t PG a) After some algebraic manipulations, this result reduces to x() = foot — fo/uilcos p(t + 7) + (a/p) sin p(t + rye“? (-1 --r and illustrates how the oscil- lator will approach the steady state of a fixed displacement as ¢ increases: é x(«0) = © = const. “ Example 2. Nondamped Oscillator, The use of the Fourier-transform and contour integration has some special features when the oscillator is not damped and obeys the DE ¥ 4+ wx = f(t). The procedure of the preceding example remains the same until we reach the expression 1 * Faye x) = du. Vor V7 APPLICATIONS OF FOURIER TRANSFORMS 283 The difference is that now the poles of the integrand are on the reat axis, namely, at w; = wo and @, = —wy. Because of this, the integral for x(¢) is not even defined until a way of circumventing these poles is prescribed. At this point an additional insight into the physical nature of the problem is desirable. For instance, the original DE may describe an actual mechanical or electrical oscillator. Such an oscillator inevitably has some degree of damping and the damping term has been omitted from the DE only because it was deemed to be small. Jf this is so, then the poles should not actually be on the real axis, but slightly below it. This suggests the rule: Replace wg by wy — ie, where ¢ is an arbitrarily small positive number. This will depress the poles below the real axis and properly define the integral for x(t), After all calculations are performed, one may let ¢ — 0 and evaluate the limit, which is then the desired result. This can be written as follows:* +2 . __ ot _ Fwy ho > ~ Te f [6 = Go = ioe = Hon — or, alternatively, ast Lf Fee ak al. ok bay” This result is, of course, equivalent to the requirement that the integral for x(1) should be taken along the path I’, as shown in Fig. 7.11 (rather than along the real axis in the rigorous sense). x) = + Path r Figure 7.18 If the problem at hand falls within the stated category, the solution is straight- forward. Sometimes, however, our DF represents some rather abstract physical osciltator and very little, if anything, is known about any conccivable damping. | In this case the following argument may be helpful: Suppose that the oscillator is at rest when 4 = —co, a situation expressed by x(— 00) = 0. Suppose, further, that the oscillator is disturbed by a rectangular pulse I/r (for 0 <1 <7), O (otherwise). fo) = { * Read the remark on p. Ut. } These two expressions are identical for any . 4 Such problems occur in quantum mechanics and in quantum field theory. 282 FOURIER TRANSFORMS V7 In this case tft iy | Le ra) = 4 [Le dt = and po * oa (een LO) = ~ 557 |. tale = wade F v9) with the path of integration as yet undetermined. ‘The integrand has two poles, at @ = wo and » = —wg, with the residues and itary give Dat 208 To simplify the argument, assume that + is very small, which implies a very short pulse. Note, however, that the height of the pulse is then very large. In this case the expression i 1— elo Tag approaches unity and the residues become* —iteg! feo emite! ot e, Ang and ~ Arup Now consider the case t < 0, Whatever the treatment of the poles, the integral for x(t) may be evaluated by closing the contour “up.” Now, from the physical point of view, the oscillator is originally at rest and is not expected to be set into motion before the pulse occurs. Therefore, we require that x(1) = 0 for 1 < 0. Jf one or both poles were to be circumvented “from below,” their residucs would contribute, and the above requirement could not be fulfilled. Evidently no other treatment of the poles except to circumvent them “from above" can be accepted. It follows then that the path of integration must again be chosen as shown in Fig. 7.11. Note: For ¢ > Othe contour cannot be closed “up” but must be closed “down.” Both poles contribute, and the final result yields x(f) = (Sin Wot)/wo (@> 0) i and represents the response of the oscillator to an instantancous impulse. * The replacement of residues by their limiting values as r — 0 is equivalent to replacing S(O) by the delta function 8(1). VW APPLICATIONS OF FOURIER TRANSFORMS 283 Remarks }. The approach outlined above may be said to be based on the Principle of causality: If one physical phenomenon causes another one, then the effect cannot be felt before the cause has occurred. If there is any reason to believe that our oscillator should obey the principle of causality,* then the above argument yields the desired result without a recourse to any damping, 2. The treatment of poles explicitly derived for an instantaneous impulse can be shown to be valid for other types of functions f(t) which vanish identically prior to some time fo (not necessarily zero) and even some others, namely those which admit solutions x{-0) = 0, Example 3. A Beam on Elastic Foundation. Con ler a beam of length L resting on an elastic foundation (Fig, 7.12); this term implies that the foundation is capable of exerting a force of reaction (per unit length) that is proportional to the displacement y(x) (Fig. 7.13).¢ The DE satisfied by the function is then af = ax) — Cy) (C = const), where 9(x) is the external force per unit length, Figure 7.12 Figure 7.13 If the beam is very long and loaded by a concentrated force (Fig. 7.12), in which case the function g(x) is reducible to P 4(x), then it may be replaced by an infinite beam, because its deflections ¥(x) are going to be very small for large values of |x|. If this approximation is accepted, then the solution can be obtained by Fourier transform methods, Using 4, 5 (| =k'YG) if 50) = Yo, * Except for some scientific speculations, the principle of causality is commonly accepted as a universal law of nature. However, the history of physics has witnessed many revolu- tionary changes and we must always bear in mind that a speculation today may be a theory tomorrow. } We assume that the beam is “glued” to the foundation so that the reaction can result in either pressure or tension. This approximates the more common case in civil engineering where the reaction can only be in an upward direction (transverse pressure on the beam). 284 = FOURIER TRANSFORMS VW we obtain (Efk* + C)¥(k) = Q(k), where O(k) = F{q(x)}. Therefore +0 ike w(x) = 5-1 YR) = + | Soe a ak, If we take® q(x) = P 8(x), then Q(k) = P/\/Tx and tag entke vO) = « Ea po* Writing El/C = a and using ak = z, we obtain +e tia we) = ae WH ® The integrand has four simple poles: zy ery eM gi a gH), For x > 0, the contour for y(x) can be closed down and only the poles zg and z4 contribute, For x 0), avi war JQ) = ~— sin a) elev? <0). x a cos ——— a2 (008-5 @ These two expressions can be combined into a single one: YQ) = ‘con lle 4 sin HL) gle gg ea (oa + Aye all x). BIBLIOGRAPHY Caastaw, H. S., Iniroduction to the Theory of Fourier’s Series and Integrals. London: Macmillan Co., 1930. SNeDvon, I. N., Fourier Transforms, New York: McGraw-Hill Book Co., 1951. Trrchmarsu, E. C., Introduction to the Theory of Fourier Integrals. New York: Oxford University Press, 1948, ' ‘Tranter, C. J., Integral Transforms in Mathematical Physics. London: Methuen & Co. Ltd., 1956. . ~ * We shall work within the space of distributions, PROBLEMS =. 285 PROBLEMS . Verify the (first) Parseval theorem for each of the three functions f(x) used in the ‘examples of Section 7.2. 2. Find the Fourier transforms of the following functions: cos Bx at xt a) SX) = ec cosfx A> 0), b) fd = coskox (|x] < Nx/ko), 109) = { 0 (xl > Nx/ke), Wis a positive integer. w Establish the following result (using the technique of Example 6, Section 2.12): {ene ax 1 cosa del <2. sinh wx) V/gq cosh k + cosa > 8) Obtain the analog of conjugation property (Section 7.3) for the case of purely imaginary f(x). b) Do the same for the case of complex f(x). 5. Establish the following formulas: -s a a) Sgfe™* cos x} = a wey’ sinx OS x }- "ie 2 2 ©) Sele") 2 wk x @+ Rp” @) Seb} = Ve Te) ZED E act, 6. Develop the formulas for dF) af) (a fo) a) sf get? b) Ss (9 ‘ 2 56 ae | Simifar to that for ¥e{d?f(x)/dx?} quoted on p. 275, Deduce the following convolution theorems: _ a) oa*(FatoG gay = —L f aU + 8) — fe - Dldt Vin Jo (where f'(x) is the “odd extension” of f(x), b) SSF AWG = x f HOI + + fe — Oa, _ ©) SSF ShGct)) = z f HEME — 8) = gle + DAE = 286 © FOURIER TRANSFORMS 8. Using the symmetry of the integrand, show that Sele} z etiite, a Let fO) = erfe 0 = O/VR) fe? at. Using the formula connecting Go {f"(x)} with Fs{/€x)} and the above result, show that 21 — eA Ssferfe Qa} = f=. 2 ' Solve the problem discussed in Example 1, Section 7.7, for the following cases: 2) overdamped oscillator, wp ? T+ | | »T fh betm | Figure 8.1 tb) * More precisely by a pair of such forces. } See Example 2 in Section 6.5. 287 288 PARTIAL DIFFERENTIAL EQUATIONS BL Now suppose that the wire can be deformed from its original, almost straight, shape by some moderate (small with respect to T) forces which are either acting permanently or are used to produce an initial displacement or initial transverse motion of the wire. It is then reasonable to expect that the stope of the deformed wire will be small everywhere during the motion (for instance, the maximum slope should not exceed 0.05). Suppose further that the wire is never subject to any other horizontal forces except the tension T, or to any initial horizontal displace- ments or velocities of its points. From all these fairly restrictive conditions we can derive certain consequences: a) The elements of the wire can have only transverse displacements, since there are no initial longitudinal displacements or motions and no subsequent net longitudinal forces. b) The tangential force in the wire is at all times and at ali points constant {toa high degree of accuracy) and equal to T. Indeed, for an arbitrary piece AB of the deformed wire the equation (see Fig, 8.2) Ty cosa = T, cos 8 must hold at all times. However, COs a = ! 2 FS El ban? ae VIF tanta Hranta {to an accuracy of 0.5% or better if |tan a] < 0.05). The same argument holds for cos 8, leading to T,=T,=T. ‘We can now derive the equation which must be satisfied by the deflection u of the wire at any point x and any instant of time #. Considering an element dx (Fig. 8.3), we obtain, noting that dx & ds within our approximations, . Tin 8 — sina) + Fae — pg dx = pdx, where p is the density of the string (so that p dx is the mass of the element), F is the external force per unit length at point x, and 47u/ar? is the acceleration of the element in the transverse direction. The above equation represents Newton’s second law applied to the element dx. ‘Within our approximations we can set fit ev tana = OH os) Vit tanta ex 2 og a cw OH ax) . sing cS (x4 4) : sina = and, similarly, aL THE STRETCHED STRING. WAVE EQUATION 289 Figure 8.2 Figure 8.3 It is then reasonable to replace sin § — sina. du/ax(x + dx/2) ~ au/ax(x — dx/2) a by the second derivative 07u/ax?, and the equation becomes vu #u Toa t+ FD — p(x)g = 0 55 * where we have taken into account that the string may not be homogeneous (so that p may vary with x) and the external force F may vary with x and ¢, This general equation for co 1) can be simplified in many cases. For instance, if p(x)g is small compared to F(x, 1), it may be neglected and the equation re- duces to a°u eu Taya t+ FO = a) oo (Note that the term (27/91) should not be neglected, because 874/31? may be large due to fast vibrations: the term T(82u/ax?) is, in general, appreciable be- cause T is large.) If there are no transverse forces ai ing on the string, then the equation assumes the form ou vu Ta — eG) aa: Finally, if p is constant, as in the case of a homogeneous string, then the equation can be written in the form 2, 2, a 72 = ® &, i ® 2 xi where c? = T/p = const (note that ¢ is real because p and T are positive). This equation is known as the (one-dimensional) wave equation; it is a linear partial differentiat equation of the second order in two variables (x and /), with constant coefficients. 290 PARTIAL DIFFERENTIAL EQUATIONS. 8.1 Although partial differential equalions are, generally speaking, more difficult to solve than ordinary differential equations, some of them have fairly simple solutions. Such is the case with the above wave equation. Consider the function ux, = f(x — ef), where f is an arbitrary twice differentiable function. It is simple to verify that u(x, f) is a solution of our wave equation. Exercise. Produce a proof of the above statement by setting z = x — ef, using ou du az ax az ax’ “I A simple reflection witl show that the function ul, 1) = g(x + ct) is also a solution of the wave equation no matter what g is.* As a matter of fact, it is possible to prove that the expression ux, = fle — et) + g(x + ct) sepresents the general solution of our PDE:t ie., any solution can be expressed in this form, Despite all this, physical problems leading to the one-dimensional wave equation do not become trivial just because the gencral solution is known. Physical problems impose certain conditions on the solution u(x, 1). For instance, since a typical stretched-string problem usually implies that the string, of a given length L, is fixed at the ends which may be placed at x = Qand x = L, it follows that the solution must satisfy the conditions (0,1)-0 and wd, = 0 for all values of the variable #. Such conditions are appropriately called the boundary conditions. Moreover, the physical problem begins at a certain instant of time, usually selected to be ¢ = 0, and it is often assumed that the initial shape and the distribu- tion of initial vetocitiest of the string are given. These so-called initial conditions are functions of the variable x and can be expressed by the statements. M0) = mols), Sx, 0) = vole), * So long as it is differentiable twice, f PDE will be used as the abbreviation for “partiat differential equation. ¥ These correspond to the familiar notions of initial position and initial velocity in Particle mechanics, 8.2 THE METHOD OF SEPARATION OF VARIABLES 291 where uo(x) and v(x) are given functions, representing the displacements and velocities of al! points of the string at the time ¢ = 0 When various such conditions are imposed on u(x, 4), then it is by no means easy (in general) to determine the functions f and g in the general solution, In many cases it is convenient to abandon altogether the form ux, 1) = £00 — ct) + a(x + ct), and to approach the problem from an entirely different angle.* In particular, we shall discuss the methods which favor compliance with boundary conditions trom the very start, 8.2 THE METHOD OF SEPARATION OF VARIABLES Since the additional conditions imposed on u(x, 1) in our string problem Fall into two groups, (a) those involving x (boundary conditions) and (b) those involving ¢ Ginitial conditions), it may be reasonable to seck solutions of the PDE in the form uG, = XO), where X is a function of x only and Tis a function of ¢ only. If X¥(x) is chosen to satisfy the conditions XO)=0, XU) = 0, then the function (x, £) will satisfy the same conditions. Then T(#) may, perhaps, be chosen to satisfy the initial conditions. ‘We now reqflire that u(x, #) satisfy the PDE, We have Puls, 2 aux, 2 ary Therefore PX, 1 yer. at 3” die Dividing both sides by ¥(a)T(1), we obtain CPx 1ider X det" FT Ge The left-hand side of this equation depends on x alone; the right-hand side de- pends on / alone. If this equality is to hold for all x and 4, it is evident that either side must be a constant (same for both sides): lax X dt * For some problems, however, this form is very suitable. 292 PARTIAL DIFFERENTIAL EQUATIONS 8.2 The constant d is known as a separation constant, The equation for X(x) can be written as ax Gee OX and will lead to exponential functions if > 0, to trigonometric functions if A < 0, and to a linear function if X = 0: Aet YE 4. Bem2V% @> 0), AGA) = YA’ cos (xR) + BY sin (xV=2) (A < 0), atx + BY Q@=0). It is not difficult to verify that the boundary conditions X() = 0, XW) = Ocan be satisfied only if » < 0 and, moreover, only if A’ is set equal to zero and the values of ) satisfy the condition V-N=a/k (n= 1,2,3,...). Exercise. Show, in detail, that it is Possible to satisfy either X(0) = 0 or XD = 0, but not both, if > 0. Also, Prove the statement made for the case < 0. These “allowed” vatues of the separation constant ), dw = —ne/E® (n= 1,2,3,...), are usually called the eigenvalues, or characteristic values, of the problem under consideration.* By this we mean the problem of finding functions satisfying the given DE and the given boundary conditions. In our case there is an infinity of such functions, called eigenfunctions, and they read Xx) = By sin (mex/L) (nw = 1,2,3,...), where B, is an arbitrary (nonzero) constant which may, in general, be different for different eigenfunctions. In our problem of the stretched string the function T(t) which is multiplied by X(x) must satisfy the DE with the same separation constant as X(x). Therefore, to each eigenfunction X,(x) there corresponds a function T,,(1) satisfying Li@m , na? OT, ae ~~ ~ Te This yields ch . axct . TA) = Cy coset + Dy sin a ; where C, and D, are arbitrary constants. * The word eigenvalue is an adaptation of the German term eigemwert derived from eigen = proper and wert = value. 8.2 THE METHOD OF SEPARATION OF VARIABLES 293 Summarizing our results we may say that the attempt to find the solution of our PDE with given boundary conditions and initial conditions in the form u(x, ) = XOOT() leads us, so far, to an infinite number of such functions which may be written as n(x, ) = (4 cos FS + B, sin me) sin, where A, = B,C, and B, = #,.D, are arbitrary constants. Each of these func- tions un(x, 4) satisfies the PDE and the boundary conditions. It remains for us to select from among these functions, adjusting the constants 4, and B,, those func- tions that will also satisfy the desired initiat conditions. Before we do this, however, note that each function uq(x, £) represents, on its own, some kind of possible mo- tion of the stretched string (corresponding to some special initial conditions). These types of motion are known as the characteristic modes (or normal modes) of vibration of the string. Each one Tepresents a harmonic motion (vibration) with the characteristic frequency (or “eigenfrequency”) @ =nec/L (n= 1,2,3,..). A bit of reflection will show that it is not possible to satisfy arbitrary initial conditions 12,0) = wots), Hx, 0) = 266%) by any single function uq(x, 4). Indeed, any particular function u(x, 1) will satisfy initial conditions of the form Chr. nex a tox) = Ay sin, pay = By E £ and 0 others. To overcome this difficulty we can lean back on the powerful principle of superposition. Our PDE is linear homogeneous, and so are our boundary conditions. [tis trivial to verify that if two functions a(x, 1) and tm(x, 1) satisly the PDE and the boundary conditions, then their linear combination IK, OO = Cyttax, 1) + Cotin(x, 1) will also satisfy the PDE and the boundary conditions. By induction this will also be true for a linear combination of a finite number of functions tg(x, 4). Itis not unrcasonable to conjecture that the same properties will hold for an infinite series formed by the functions u_(x, 1): . nxct a Rch\ Nex YX, = x (Ancos ae +B, sin sin os. provided the series converges (or if not, provided it can be treated asa distribution, in conformity with the principles stated in Chapter 6). 294 © PARTIAL DIFFERENTIAL EQUATIONS 82 To be more precise, the function defined above certainly satisfies the boundary conditions 70, = WL = 0 (since each term satisfies them). It will also satisfy the PDE, provided it can be differentiated twice term by term with respect to both x and ¢. From the point of view of the classical theory of functions, this may impose considerable restrictions on A, and By; €.g., the twice differentiated series p(x, 1) must be uniformly con- vergent, On the other hand, from the point of view of the theory of distributions it is sufficient* that A, and B, are of the order of some power of 1: |4n| < An”, |B} < Bn” (some 4), and this point of view may be adequate for physical applications. In any case, however, the justification of the method depends on the properties of A, and B,, and it is usually convenient to assume its validity and justify it later, after the solution has been developed. It is needless to emphasize that the function y(x, 1) is a Fourier sine series in x (it is also a Fourier series in #, but this fact is of much less importance). Setting t = Oand using the first initial condition, we obtain uo(x) = 5 A, sin a. at If the function wo(x) can be expanded into a Fourier sine series, the coefficients A, can be determined. In physical problems #9(x) is invariably continuous, piece- wise very smooth, and vanishes at x = O and x = L. Therefore, it can be repre- sented as above. Similarly, calculating (9u/2t)(x, ¢) and using the second initial condition, we can obtain cae. frx vo(x) = Ea, Ein a In physical problems v9(x) is sometimes assumed to be discontinuous. However, it is invariably piecewise continuous and piecewise very smooth, and the coefficients B, can be determined as well. Consequently, we have constructed a solution to our problem in the form of a series wo 0= Ye (4 cos + By sin") sin "=, i which satisfies the boundary conditions and initial conditions for all physically reasonable functions ug(x) and vo(x). * See p. 253, 8.3 LAPLACE AND POISSON EQUATIONS 295 8.3 LAPLACE AND POISSON EQUATIONS Many problems in electricity and magnetism are also reducible to partial differen- tial equations. For instance, one of Maxwell’s equations reads (in MKSA units) ou curlE Hog, = 9 In the so-called electrostatic case, where the magnetic field H is absent (H constant in time leads to the same situation), this equation reduces to cule = 0. Since this holds everywhere in space it is possible to derive the electrostatic field from a scalar (the electrostatic potential),* E = ~grady. Hf this property is combined with Maxwell's first equation (for vacuum), divE = p/eg, then we obtain div grad p = Vp = —p/ep, where p is the charge density (assumed to be given as a function of coordinates) and €, is constant (electric permittivity of free space). An equation of the type (r = position vector) Wo = f(r) is known as the Poisson equation. Its particular case, when f() = 0, Vp = 0 is known as the Laplace equation. As follows from the above, it must be satisfied by the electrostatic potential g(r} in a charge-free region. Example. A very long, hollow, metat cytinder of radius a is split lengthwise into two halves, insulated from each other. The halves of the cylinder are kept at potentials +V and — V, respectively. It is desired to find the electrostatic ficld E inside the cylinder. This problem may be solved (approximately) by an idealized infinitely long and infinitely thin split cylinder (Fig. 8.4). It is convenient (see the boundary conditions below) to treat the problem in cylindrical coordinates, as shown. * Minus sign is conventional, 296 PARTIAL DIFFERENTIAL EQUATIONS 8.3 The field E is easily found from the electrostatic potential g which must satisfy the Laplace equation written in cylindrical coordinates laf ag La%y , a% 1 2(-22) 4 bee et = % From symmetry considerations it is evident that » must be independent of z, so that o = off, 8). The following boundary conditions are imposed on y: Aa,A=+V ( 0), and C = const and log r (for # = 0). Since g(r, 8) must possess a gradient at r = 0, the solutions 7 and log r are not admissible. According to the general principle, we form a linear (infinite) superposition of solutions: #6, 0) = Ag+ 35 r"(Ay cos m6 + Bp sin mn). . ean 8.4 THE DIFFUSION EQUATION —-297 To satisfy the boundary condition nm ITY O< £(5) sin mé. m=1,3,5,... 7 It is not difficult to verify that this function satisfies the conditions of the problem. In particular, if r < 1, where r, is any positive number less than a, the series can be differentiated term by term twice with respect to both r and 6 (the factor r/a will cause uniform convergence of the twice-diflerentiated series), Thus the PDE is easily verified. Remark. From a physicist’s point of view, o(r, 8) does not have to salisty the PDE at + = a, so tong as it approaches the prescribed boundary conditions [which it does since at r = a it reduces to the Fourier series for f(O)). 84 THE DIFFUSION EQUATION Another type of PDE arises in the study of diffusion of some substances through @ continuous medium, such as diffusion of smoke through the air, diffusion of neutrons in a nuclear reactor, or diffusion of a dissolving chemical through the solvent. Let p denote the density of the substance (often called the concentration) measured in units: “stuff"/cm*, where “stuff” is the appropriate unit for the substance, such as the number of neutrons or the number of atoms in the solute. Let j denote the current density measured in “stulf”’/cm?2 - sec (amount of sub- stance passing per unit time through a unit area normal to the direction of flow). 298 PARTIAL DIFFERENTIAL EQUATIONS 84 Then most diffusion phenomena obey the following linear law, sometimes called Fick's law: j= ~Dgradp, where D = const is the diffusion coefficient, which depends on the properties of the medium (D is measured in cm?/sec). Suppose, furthermore, that the diffusing substance is neither absorbed nor emitted by the medium. Then the equation of continuity (conservation law for the substance) 8s dvi #4 dvj=0 must hold. Combining the above two equations, we obtain. % _ ppt $x Dv’. This result is known as the diffusion equation. An immediate modification is possible if the substance is capable of being absorbed (destroyed) or emitted (created) as is the case of a chemical compound or neutrons (absorption and emission by the nuciei). In this case, the equation of continuity reads 2 diy = 3 tdivi= 5, where s is the source density, i.e., the amount of substance created (or destroyed, if s < 0) per unit volume per unit time. The expression above gives rise to the nonhomogeneous diffusion equation op 2, 3p = DV +s. Remark. In some problems, one encounters the steady state of diffusion processes. Then is independent of time and the equation reduces to the Poisson (5 x 0) or the Laplace {s = 0) equation. Some other physical phenomena, such as conduction of heat, lead to the same type of equation. Conduction of heat is usually assumed to obey the linear heat flow equation (first established by Fourier) a= —k grad u, i ! where q is the heat current density (cal/em? - sec), w is the temperature (deg), and k is the thermal conductivity (cal/cm - sec- deg). For an arbitrary volume V enclosed by a surface S the heat influx per unit time is — $s (qd) (the minus sign is traced to the conventional definition of dS to be directed outward), 8.5 USE OF FOURIER AND LAPLACE TRANSFORMS 299 If, in addition, heat is generated (e.g., by an exothermal chemical reaction) at a rate given by s(x, y, z) per unit time and unit volume, then the total heat received by V in a time interval A¢ is Q= [-ffv-as) + fffsav]a According to the basic formula of calorimetry, this amount of heat will raise the temperature within VY by an amount Au so that Q= Audv, Ife ue where p is the mass density (gm/cm?), ¢ is the heat capacity (cal/gm - deg). (It is tacitly assumed that V is sufficiently small and a unique temperature u may be assigned to the entire volume.) Thus we are led to the statement ffor-0 = |ff [ood — Jan If At is very small then Au/Ar may be replaced by du/at. Applying the divergence theorem,” we deduce that — diva ws Ou Combining this with the heat flow equation, we obtain = = aVu + bs, 3 a cen), b= (25). cp \ sec cp cal where In particular, if no heat sources are present, this result reduces to du 2y2, a 7 V%n, which is often called the heat-conduction equation. $5 USE OF FOURIER AND LAPLACE TRANSFORMS Example 1. A very long and narrow pipe is filled with water. At time f= Oa quantity of salt (Af gm) is introduced into the pipe at some point x9 (remote from both ends of the pipe). We wish to find the concentration of salt at any later time. * See Section 1.8. 300 PARTIAL DIFFERENTIAL EQUATIONS. 85 We can idealize this physical problem by assuming an infinitely long pipe (Fig. 8.5). Moreover, since the pipe is narrow, the variations of concentration of salt over a cross section of the pipe may be neglected and the concentration can be treated as a function of two variables only, x and #: P= p(x!) {in gm/cm*), This function satisfies the one-dimensional diffusion equation pie, of ax? The initial conditions can be idealized by the statement that p(x, 0) is zero every- where except at the cross section x ~ Xo, where M gm of salt have been intro- duced. This statement can be expressed by means of the Dirac delta function (x, 0) = (Af/A) 6x — x9), where A is the cross-sectional area of the pipe. The boundary conditions for this problem may be formulated by the statement fim Pt, = p(to,9=9 (alls), reflecting the physical law of conservation of the total amount of salt. Salt. introduced at 1-0 x Je-~ -x9—- Figure 8.5 Exercise. Show that if p(x, 1) > po > 0 for all x > N and some ¢ (some fixed po and NY), then the total amount of salt cannot be finite. The boundary conditions of this problem bear a certain resemblance to the problem of the vibrating string in Section 8.2. The basic difference is, however, that the sought furiction p(x, ¢) must now vanish at the ends of an infinite intervat {rather than at points x = 0 and x = L). This fact suggests that the solution is more likely to be expressible by means of a Fourier transform than by a Fourier series. As a matter of fact, since p(x, 1) must be very smooth with respect to x, that is, it must possess a second derivative, and since it is absolutely integrable (bear in mind that p(x, t) > 0), t [0 0 0A dx = M, it follows that it must possess the Fourier transform, 85 USE OF FOURIER AND LAPLACE TRANSFORMS. 301 40 R(k, ) = F{olx, O} = [ atx, Ne dx. Vix We shall assume the vatidity of differentiation under the integral sign,* namely, ARK) Lf ape ite gy dt ix Jw “Ot as well as the usual property 5 {2%p/ax7} = —K2R(k, t). Then R(K, 1) satisfies the ordinary DE : anh 8 5) - _ DeRK, 1, and is given by : Rik, 1) = Rlk, NeW PP Given the initial condition, it now follows that sh eto RK, 0) = o {i ocx ~ xo = u = {in the sense of the theory of distributions). Therefore M _ itkeoDkY R(k, 1) = ——x- ee . AVde ‘The inversion is not difficult (see, for example, p. 231), and the result is Mod (229 AE p=. Le A Vay Dt This should then be the solution to our problem. itis easy to verify that it satisfies the PDE and the boundary conditions (either as a distribution or as a “classical” function). Also, lim p(x, ) = (M/A) B(x — 2x0), in the sense of distributions, of course, This is all that is required to adopt p(x, 1) as the solution of the physical problem. If desired, the steps taken in the Fourier transform technique can also be justified. * If we postulate that a(x, #) is a distribution (see p. 242) with respect to both x and 7, then this procedure can be shown to be automatically valid (by methods similar to those in Section 6.9). Otherwise, the idea mentioned on p, 294 can be invoked. A justification a posteriori will be possible after p(x, ) is found. 302 PARTIAL DIFFERENTIAL EQUATIONS 8.5 Bxercite. Show that (in the sense of classical functions) for any ¢ sé 0, p(x, 1) is ex: pressible by te ete) = mz } Rk Ne at and that it can be differentiated twice under the integral sign with respect to x, and once with respect to 1. The sketch of the obtained solution is given in Fig. 8.6.* ol) Smalt ¢ Figure 8.6 Example 2. A very long and narrow cylindrical rod thermally insulated over its lateral surface is kept at zero temperature up to ¢ = 0 at which time one of its ends is placed in thermal contact with a heat reservoir at the temperature uy. It is desired to find the temperature in the rod. at any later time. 110. = ae Figure 8.7 As in the preceding example, the temperature in the rod can be treated as a function of only two variables, w= u(x, 1), and the rod can be’ considered to be semi-infinite (Fig. 8.7). The problem may be mathematically formulated as follows: u(x, £) must satisfy the PDE. du _ 9a’ ou ou ar xt with the initial condition u(x, 0) = Oand the boundary conditions 40, =u, — u(oo, )= fim u(x, ) = 0. * Different scales are used for the ordinates; otherwise, the area under each curve should be the same. 85 USE OF FOURIER AND LAPLACE TRANSFORMS = 303 This problem can be conveniently solved by the method of Laplace transforms. Indeed, from the second law of thermodynamics, it follows that, (Xx, 1) S to} therefore u(x, () must possess a Laplace transform. It is reasonable to assume that its time derivatives also possess a Laplace transform (except at ¢ = 0)so that if UG, 3) = L{uG, OF, then PU, 3) _ s Ge) 5 us 9), which is obtained by taking the Laplace transform of the PDE and using the initial condition. Solving (with boundary conditions applied), we find that UG, s) = UO, se = * ent als If the available table of Laplace transforms does not list this function directly, then the rule £ { [ “fe ar} = 1h), where F{s) = £{f(}, may, perhaps, be used. If even this is not possible, that is, F(s) = 7" is not listed, then we may employ Mellin’s inversion formula vee uo st, valo yas = ds. u(x, ) wo fo bee fs, Bearing in mind the branch cut created by 4/s as well as the contribution from the origin, we close the contour as shown in Fig. 8.8. Lmploying an argument similar to that in Example 4, Section 2.15 or, alterna- tively, setting s = fo and making use of Jordan's lemma, we see that the contribu- tion from Cg vanishes in the limit R — a0. In the contribution from 4 to B we may set Tm = oe > oe™, and in the contribution from C to D, we may set s= oe oe, 304 PARTIAL DIFFERENTIAL EQUATIONS 8.6 where ¢ is real and where the limits 6 45 are anticipated. The sum of these contributions reduces ta wo f ent sin XE dor. ® o a Write /o/a = k and reduce this to 2ug f * (neti sin kx dh x Jo “e This integral may be evaluated, ¢.g. from the formula f oO cos wt du = Lk ay 0) o a by integration over w from 0 to x: f dof eo cos wu du ° a Therefore The contribution from C, yields —u. Therefore Vie M st—2¥ija ds wD af ees x x = = (0 + uo rf) = uy erte 2. ( a ) a lat 8.6 THE METHOD OF EIGENFUNCTION EXPANSIONS AND FINITE TRANSFORMS The selection of methads used to solve physical problems involving partial differ- ential equations is based on the geometry of the problem and the character of the boundary conditions. For instance, the choice between Fourier transforms and Fourier series (atising from the separation of variables) can be traced to the Fact that in one case the Tange of variation of x is infinite, that is from —oo to +o, and in the second case it is finite, from 0 to L. In many respects, however, the method of Separation of variables and the method of transforms are closely related, To illustrate this, let us consider again the problem of the vibrating string solved in Section 8.2. 86 HIGENFUNCTION EXPANSIONS AND FINITE TRANSFORMS 305 This problem consists of finding a sotution 4(x, 1) of the one-dimensional wave equation o_o ox? ~ oF ar satisfying the boundary conditions x(0, 1) = ul, 1) = 0, and the initial conditions #0) = wal), 46,0) = vex). Instead of looking for solutions of the PDE as we have done in Section 8.2, let us focus our attention on the boundary conditions. Our solution u(x,t) is evidently a very smooth function with respect to x and vanishes at the ends of the interval (0, £), Therefore, it must be expressible in a Fourier sine series appro- priate for this interval: ux, ) = = anCi) sin 2 It is worthwhile to emphasize that this statement is completely independent of the fact that the functions sin (nwx/L) “have something to do” with the solutions of the wave equation vanishing at x = 0,L. True enough, this fact is of great help later, but, in principle, the above formula must be valid no matter what Kind of PDE the function u(x, 4) satisfies. Let us now differentiate u(x, 1): O74 So nen? Led axe ~~ xy “Ex bal) sin “>> au zs aby. ex a” de Substituting into the PDE, we obtain sy {1 d’bn nix? - eX This equation must be valid for all x, implying that the coefficients with each sin (nxx/L} (x = 1, 2,3,. ..) must vanish: L dbp nx? @ de pr n= 0. In case this conclusion docs not sound rigorous, the following argument may be invoked: Multiply the series by sin (mxx/L) (fixed mm) and integrate over (0, L). 306 = PARTIAL DIFFERENTIAL EQUATIONS 8.6 Assuming the validity of term-by-term integration® and using the orthogonality property L «RX. ranx f sin sin —-~dx = 0 (a # m), we conclude that only the term with a = m survives and the DE for &,(t) follows. It is now a simple matter to solve the DE for 4, and obtain nret +. Axct BO = An cos + &, sin rc The sought solution u(x, 1) is therefore u(x, 1) = x (4 cos Ft. By sin me) sin, a and the coefficients A, and B,, are obtained from u(x) and vo(x) in the usual way: of | ex An =F f u(x) sin 7 x, 2 f® - RX By = af B(x) sin ~~ dx. This type of approach may be called the method of eigenfunction expansions, since we start by expanding u(x,t) in a series of some suitable eigenfunctions. Tecan be cast in a form similar to that of the Fourier transform approach: This time we again start with the PDE au 1 ae ox? 7 2 ot Let us multiply this equation by sin {nrx/L) and integrate over x from 0 to L: b b Ou se mex ot fae nex _ f gah pdt = 5 fo a My ax @ = 1,2,3,.. ‘This procedure is analogous to the method of taking the Fourier sine transform with one difference: the “wave number” k in sin kx is not arbitrary but restricted to discrete values k = nx/Z, and the range of integration is a finite interval (these two facts are intimately related). It is from this Point of view that we may talk about applying the finite sine transform to the PDE. * From now on we shall tend to omit mathematical details of this type; see p. 294 for a general approach to such problems, 86 EIGENFUNCTION EXPANSIONS AND FINITE TRANSFORMS = 307 The right-hand side of our equation can be written as Bex 2 Le ux, a) sin “dx [with appropriate assumptions about 1(x, 1]. As for the left-hand side, we inte- grate by parts: bo L [bem i Seon ‘9 Ox? L ex QL nef" au nex otk lo axe’ EE ax * ou aex “TT dy ix [by virtue of the properties of sin (xx/L} at the boundary}. It is at this point thal the choice’ of the “transform by sin (nxx/L)" is to be appreciated. Let us integrate by parts once more: Bg gh -oE f asin = dx ° nex "2 oe a = ™ neos L Ele 2 2a fh --22 [ usin de, 0 The net result of these derivations is the ordinary DE Ee Le t a nx’ 2 ax ba f wen Din ae Ef x, 1) si 7 dx for the function L fn,0 f ux, )sin "de (a= 12,3... ° In accordance with the previous notation this function will be denoted by Fe, ) = Fb) so that 5,(/) satisfies Pot) _ as me de bl) (2 = 1,2,3,...) and the remainder of the problem is as before. Remarks 1, It is to be firmly remembered that the function &,(f) can be considered as a repre- sentation of the function u(x, t): If 6.(t) is given, then u(x, 1) is determined (by summing the Fourier sine series); the function 6. is a function of two variables; one continuous (¢) and the other discrete (1). 308 PARTIAL DIFFERENTIAL EQUATIONS 87 2, The function 6,(1) is a representation of u(x, 1) with respect to a particular set of functions (sin (mrx/L}}. The success of the method is based on three facts: (a) the fune- tions sin (vxx/L) satisty the same boundary conditions as u(x, , (b) the functions sin (wxx/L) ace orthogonal to each other, and (c) the PDE is significantly simplified for the chosen représentation: it becomes an ordinary DE and of a simple type 3. The method of eigenfunction expansion and the method of finite transform are essentially identical. They are similar to the method of infinite transforma. Which partiow, lar approach is more useful in a given problem depends on the geometry and the nature of the boundary conditions. The method of separation of variables is then a general tech- nique which Teads us to the particular eigenfunction expansion or transform appropriate for the problem at hand. 8.7, CONTINUOUS EIGENVALUE SPECTRUM The method of separation of variables applied in Section 8.2 to the PDE. along with the boundary conditions uO, ) = u(L, 1) = ©, has led us to a set of eigenvalues wy? re — F @ = 1,2,3,...) corresponding to the eigenfunctions Xn(x) = Bi sin (nxx/L) (un = 1,2,3,...9 of the separated equation xX ae This set of eigenvalues is commonly called the Spectrum of the differential operator a ae which is regarded as being applied to the functions f(x) satisfying the boundary conditions FO) = f(D = 0. Among ail possible functions satisfying these boundary conditions, the particular functions ‘ %Q(x) = sin (tex/E) (8 = 1,23..), satisfying X"’ = XX, are called the eigenfunctions of the differential operator 0. * Compare the above with the idea described on b. 187, RT CONTINUOUS FIGENVALUE SPECTRUM 309 From the preceding section, it follows that these functions play an important part in the problem of construction of solutions of the wave equation. This spectrum happens to be discrete as is usually the case with the boundary conditions on a finite interval, However, if the interval is infinite, the eigenvalue spectrum (of the same operator ©) is usually continuous. Remark, The term spectrum is, of course, borrowed from optics (so is the symbol )). The quantum-mechanical description of an atom invoives an eigenvalue problem and the resulting eigenvalues are related to the wavelengths of light waves emitted by the atom. Example. Four large conducting plates are arranged as shown in Fig. 8.9 with their respective electrostatic potentials. Since the size of the plates is much larger than the separation 2a, they can be treated as though they were extended to infinity in the x- and z-directions. We wish to find the electrostatic potential in the region between the plates. We find that the mathematical formulation of the problem requires the electrostatic potential (x, y) to satisfy ae ge tae 7? and the boundary conditions o(x, 4) = ox, -a)=+V (> 0), 0,4) = 9, -Q=-¥ (<0) Fignre 8.9 or, more compactly, g(x, ba} = F[2S(x) — 1], where S(x) is the step function. Let us apply the method of separation of variables: If p(x, y) = X()YQ), then PX fy _ aa Ga In order to determine the eigenvalues, we should use the boundary conditions. The boundary conditions apparently involve the variable y, but they do not seem to favor any particular choice of }. It appears that all three possible forms AY, YQ) = A, cos(VAy) + By sin(Vay) = (A> 0), @ YO) = Azt By (= 0), e ¥Q) = Ase" + Bee A <0) Q) could, conceivably, form a linear combination satisfying the boundary conditions. The reason for this indeterminacy is not hard to find: The boundary conditions with respect to y are nonhomogeneous, namely, they do not demand that g(x, a) and (x, —a) vanish but rather that they reduce to a function of x, namely, 310 PARTIAL DIFFERENTIAL EQUATIONS, 87 SC) = VA2S(x) — 1]. Compare this with the vibrating string problem of Sec- tion 8.2, The requirement that y{x, a) = 0 (this is a homogeneous boundary condition) could have been satisfied by requiring that ¥\(-ta) = 0 [for cach cigen- function Y\(y) corresponding to an eigenvalue XJ. However, there is no simple condition for ¥,(p) such that (x, =a) will be made equal to some f(x). A way out of this difficulty may be found if we recognize that there are also certain boundary conditions involving x which we have missed so far. An important property of the boundary conditions is that they are antisym- metric (add) with respect to x: o(—x, £4) = —o(%, +a). This fact suggests the possibility that the entire solution is antisymmetric with respect to x: o-% Y) = —¥(%, ¥)- That this must indeed be true follows from physical considerations.* In particular, the antisymmetry implies that 00, yy=0 (-a 0). Furth¢r- more, the condition (0, y) = 0 rules out the function cos kx, Therefore XG) = Apsinkx (k > 0). . * Compare with Section 8.3. 87 CONTINUOUS FIGENVALUE SPECTRUM = 311 There is no further restriction on k and we conclude that the eigenvalues \ of the problem, X" =zhX, X0)=0, Xo) < mw, form a continuous spectrum (—oo < <0). There is a nondenumerable set of eigenfunctions X.(x). Now consider the equation for ¥(y). Since is bound to be negative Q = —k), we obtain Ya) = Bye” + Cre". Having learned the importance of symmetry considerations, we can also recognize the fact that the solution (x, y) must be symmetric (even) with respect to y: o(%, —y¥) = ey) (all x). ‘We shall impose this condition on each Ye(y) which implies that By = C), or, in other words, ¥4Q) = Dy cosh Ky. Remark. This requirement would, in any event, follow from further analysis, but this step simplifies our task. We shall now attempt to solve the problem by constructing a “linear combina- tion” of functions pul, ») = Agcoshkysinkx —(k > 0), (where the constant. Dy has been absorbed into A,). Instead of an infinite series, as was the case with the discrete spectrum (infinite but denumerable), it is reason- able to expect that g(x, y) should be expressible as an integral over a continuous (infinite nondenumerable) spectrum of eigenvalues. Indeed, an integral can be regarded as a form of linear superposition, and the analogy between the Fourier series and the Fourier transform strongly indicates such a possibility. Thus we conjecture that eG, ¥) = f Ay cosh ky sin kx dk. This is, of course, equivalent to the statement that g(x, y) is expressible in terms of its Fourier sine transform, namely, ox, 9) = V2/nf” ae, y) sin kx dk, 0 where: &(k, y) = Vx/2 Ay cosh ky, It remains for us to evaluate Ay. Since & is a continuous variable, it is more appropriate to replace A, by A(K). This function can be determined from the requirement . (xa) V = f * (ky cosh ky sinkxdk — (k > 0). 312 PARTIAL DIFFERENTIAL EQUATIONS 87 Note that once this is satisfied, all other three original boundary conditions will be satisfied because of the symmetry properties of the functions sin kx and cosh kp. According to the inversion formula for the Fourier sine transform, we have 4k, ») = V2]e { fx, y) sin kx dx, 0 In particular, for y = a, &(k, a) = Vx/2 Alk) cosh ka — V2/af” Vsin kx de. D The needed integral ‘ f ” sin kx dx 0 is divergent. Nevertheless, it can be treated as a distribution, in which case,* fF sinkx dx = 1k (k> 0). 0 Therefore and 2Vv f° cosh ky sin kx etx, y) = % coshka kM It is not hard to see that we could have arrived at the same result by applying the Fourier sine transform directly to the PDE. Indeed, let us assume that y(x, ) possesses the Fourier sine transform with respect to x, along with its derivatives dp/ax and a%—/ax*. Then Sslelx Y= Oy), as (aes aes, 2 ~ K(k»). and ax? Therefore, the transformed PDE. reads — Kalk, y) + = Solving and applying the requirement that #(K, y) is symmetric with respect "9 y [because (x, y) is so}, we obtain (k, y) = Ck) cosh ky. * The sequence fy(k) = ff sin kx dx = (1 — cos kN)/k divergesas N— «but itis weak- ly convergent for a suitably chosen set of test functions g(k) defined for 0 < k < @. 88 VIBRATIONS OF A MEMBRANE. DEGENERACY 313 The function C(4) can be obtained from the condition Hk, a) = FV) = V2/m (V/k) and the remaining procedure is the same as before. It is still necessary to prove that the obtained function kx p coshka 2v f° cosh ky ots, = ak is indeed a solution of our problem (we have employed some conjectures and assumptions), First of all, note that the integral converges: For |p| < a, we have cosh ky cosh ka ” sin kx [2a certainly exists (it is equal to x/2). Next, it is evident that g(x, y) is symmetric in y, antisymmetric in x and reduces to V for y = a and x > 0. Finally, we must prove that g(x, y) satisfies the Laplace DE. If p(x, y) can be differentiated under the integral sign, this becomes obvious. Differentiation is possible, provided bt < 4, since one can easily prove that pn cosh ky fimk cosh ka — EG 950) = vole»? * Note that lhe “free boundary” must still be able to exert a horizontal tension T. 316 © PARTIAL DIFFERENTIAL EQUATIONS 88 Example. A rectangular membrane with sides a and 6 is clamped at the boundary. It is desired to analyze the motion u = 1x, yi 1) in terms of the characteristic modes of vibration. Set up a cartesian coordinate system with the axes along two edges of the membrane. The boundary conditions assume the form (0, ¥3 1) = ula, ys 1) = 0, u(x, 0; 4) = u(x, B51) = 0. Set u(x, y3) = X@)YQ)I() and, separating the variables in the wave equation, obtain 2. a) o = eT a b) Gat ae PY (with the provision that Ay -F A2 = )- cs) wpe AY From the boundary conditions it follows* that 202 22 a Em 23,009 ee LTD leading to a double-Jabeled set of cigenfunctions Yn, 9) = Xn(X) ¥aQ) = sin uy, THER sin MEY @ 6 with the corresponding “time factors” Ton) = Arn COS Omat + Brin Sitt nant, toma = rev t/a? + 2/62 (mat = 1,2,3,...). Each pair of integers (7, 7) corresponds to a particular characteristic mode of vibration of the membrane. A particular solution u(x, 3; 1) is then sought as a superposition of all modes: where ul, 950) = SE Aen £08 unl + Br 80 Sng’) sin MEX gin « wat te ‘The right-hand side represents, of course, a double Fourier sine series, The coeff cients Amn and Byx can be obtained in the usual fashion (see Chapter 4). ; Set t= 0; then ' Holey) = De Ann sin EE sin A . mola * Compare with Section 8.2. 88 VIBRATIONS OF A MEMBRANE. DEGENERACY 37 Muttiply by sin (m'xx/a) sin (n’xy/b) and integrate over x from 0 to a and over y from 0 to & (in other words, take the double integral over the area of the membrane). Assuming the validity of term-by-term integration, conclude that the only term on the right-hand side which will yield a nonvanishing integral is the term with m =m” and » = n', due to the orthogonality of our eigenfunctions. This leads to the formula a fb _4 . Mixx. ney Ann = 4f f u(x, y) sin a sia dx dy. In the same fashion, evaluating the series for du/at and using vo(z, y), obtain - _ 4 | MRK Try Bon = ghar [ f vol ») sin EE sin BP ctx dy. ‘The assumptions necessary for the validity of all these operations are invariably satisfied (at least in the sense of the theory of distributions) in all physical problems. Our solution is, therefore, complete. Let us now look into some of the features of characteristic (or normal) modes. If the membrane vibrates in one of its normal modes, then all its points will execute harmonic motion with frequency wen? lx, 951) = Cyan sin 2% sin "FP c08 (onnt + Sun) a b (for convenience, we have set Amn = Cm C08 dun and Ban = —Can Sint Snn)- In this motion some points (x, y) of the membrane will remain at rest'at ali times t, for instance, the points with x = a/m {and arbitrary y). Evidently, such points will lie along certain curves, called nodal curves {in analogy to the nodal points for a stretched string). (2, j= mede (, 2) - mode _— Nodal lines (a) «o) Figure 8.13 As a particular example, consider the (2, t)-mode (that is, w= 2,” = 1). ‘The eigenfunction reads = sin 2® sin 2 - Voix, y) = sin sin F The only nodal curve is the straight fine x = a/2 (Fig. 8.13a). Similarly, the (1, 2}-mode ( = 1,2 = 2} has the nodal line y = 6/2 (Fig, 8.136). 318 PARTIAL DIFFERENTIAL EQUATIONS 8.8 Generally speaking, cach mode vibrates with its own frequency Wm, given by Onn = crv mia? + 12/82. However, sometimes two or more modes possess the same frequency. This is possible only if a/6 is a rational number. As an example, consider the case of a square membrane (a = 5). The €2, 1)- and (1, 2)-modes will vibrate with the same frequency {namely (cx4/5/a)]. It is then said that this frequency exhibits de- generacy (one may regard wn a8 an eigenvalue of our problem; the eigenvalue is degenerate if there is more than one eigenfunction corresponding to it). Furthermore, it is evident that any linear combination of (2, 1)- and (1, 2)-modes will represent a harmonic motion of the membrane (of the same frequency). For instance, the motions given by nis ys) = (A4sin2# sin 7 sin ®® 4. A sin ™ sin 2D) cos SY orV5 lx, YD) = (4 sin 2X sin Z sin? — Asin ™ sin = sin 22) cos a5, Tepresent harmonic vibrations of the membrane with nodal lines as shown in Fig. 8.14. These solutions are often called Aybrid modes. From the physical point of view they must be regarded as characteristic modes in their own right and may be treated as basic solutions for this particular frequency {the original (2, 1)- and (1, 2}modes are linear combinations of w; and 2]. tabs ¥3 ual, V3 spe lines |, Figure 8.14 Generally speaking, if a frequency is degenerate, there is an infinite number of associated modes, formed by all possible linear combinations of basic (linearly independent) modes, Some of these modes possess nodal curves which are not straight lines, Remark, Yn general, hybrid modes are not orthogonal to each other. In most cases it is convenient to work with orthogonal modes, and because of this some may be préfer- able to others. Exercise. Show that the modes #1 and uz given above are orthogonal to each other, Construct a pair of nonorthogonal hybrid modes. 89 PROPAGATION OF SOUND. HELMHOLTZ EQUATION = 319 8.9 PROPAGATION OF SOUND. HELMHOLTZ EQUATION Experiments indicate that the phenomenon of sound in air is related to rapid oscillations of air density and pressure about some average values pg and po. These changes are caused by the so-called collective oscillatory motion of air mole- cules described by their macroscopic velocities.* Consider a small quantity of air within a volume V (at some particular instant £) and let ¥ ~ v(7, #) be the macroscopic velocity of air at point r and time #. If ? = a(t, #) is the density of air, then the force per unit volume is given by Newton's second law: a {1.90 = 97 The total force on volume V is then = [[foger If the weight and other body forces, namely, those depending on the volume, are negligible, then this total force is equal to the surface integral of pressure (vis- cosity is neglected as well): [fae- - fpous. Transform the surface integral by means of the formula from vector calculus, + ff pas = fffss pav, and use the fact that ¥ is arbitrary to obtain the equation of motion in differential form: ptt + grad p - 0. To this formula we may add the equation of continuity (conservation of matter): 5 flo - fhe [for —f * As distinct from their random thermal motion. Usually, the thermal motions almost “cancel out” on the average. The statistical net velocity of a macroscopically small (but still containing millions of molecules) group of adjacent molecules is called the macro- scopic velocity. { This formula follows from the representation of the gradient on p. 33. 320 PARTIAL DIFFERENTIAL EQUATIONS 89 By use of the divergence theorem and the arbitrariness of V, we deduce that ap rr + div (or) = 0. Finally, the pressure p — p(r, 1) and the density p — p(t, #) can be related by thermodynamical formulas. For the frequencies involved in sound waves (20- 20,000 cps), the osciflations of p and p take place so rapidly that there is no time for an appreciable amount of heat to be exchanged between two adjacent volumes of air during an oscillation. In other words, the expansion and contraction of air are adiabatic and Bua(VN L(2Y yas, Po \Vo/ ~ \po. “ey where c, is the specific heat at constant pressure and cy is the specific heat at constant volume. The propagation of sound is governed, therefore, by the fol- Jowing equations: a ap _ _ ey eg tenadp=0, a tdiv@n=0, p= py bo, Remark. Since ¥ is a function of both position and time, the derivative d¥/dt is the total or hydrodynamic derivative” dy av fen ag t Brad) A considerable simplification of these equations is possible for a typical acoustic wave of moderate intensity. Consider, for instance, a sound wave of frequency 500 sec~", where the air particles vibrate with an amplitude of the order of 10—S cm. The wavelengtht is approximately 3.3 x 10 cm/sec 500 sec—t 60cm, where c is the speed of sound and fis the frequency. The speed v of air particles is of the order of Dw 2xfa = 2x X 500 X 10-8 ~ 3X 107! see“! * Consult any textbook on fluid mechanics; =(344246,2)edguite. (+ grad) ¥= (uf test ag) (ed + oi + edd. + For the purpose of general orientation, take the example of a plane harmonic wave with ks = acos 2n(x/d — fi), where w, is the displacement of air particles. 8.9 PROPAGATION OF SOUND, IIELMHOLTZ EQUATION 321 The velocity gradient is of the order of mw 2eX3X 107% 4 scent ay ~3X 10“ sec,” while the rate of change of v at a given point is of the order of [3 we Info & Ix X 500% 3 X 10? ~ 10.¢m/sec?, The term dv/a¢ is clearly much larger than the term (¥- grad) v, the latter being of the order of lev grad) o| ~ @ ~3X 107 x 3X 107% ~ 10“ cm/sec? so that dy/dt can be replaced by 3+/at. Moreover, the relative changes in density are quite smali: If a column of air d/4 in length is compressed by an amount a = 10cm, the increase in density is of the order of =P a _ totem ¢ = 9/4 13cm (2) =CU+ ay ely rv and p & po(l + ¥8), implying that the relative changes in pressure are also small. Finally, in the expfession 1077. At follows that div (pv) = 1- grad p + pdivy, the first term on the right-hand side of the equation is smaller than the second one by something like # factor of 107 (since p ~ 10—* gm/em’, |div v] ~ 3 X 10-* seo, [grad p| < 407?" gm/em4). Therefore we are left with a simplified (but stil! quite accurate) system of equations: oe di pony + grad p ~ 0, 3,40 ivy = 0, P= poll +¥5), =P — PO, Po where p hag been replaced by po, in accordance with §~ 10-7. From the last equation it follows that grad p = poy grad 8 so that {also use dp/dt = po (03/8t)] ay Pot = 8b divy = at po Bale 0, 3; + divy = 0. 322 PARTIAL DIFFERENTIAL EQUATIONS 8.9 Taking the divergence of the first equation and the time derivative of the second and combining, we obtain 28 ey% or > where c? = po¥/oo. The relative change in density & obeys, therefore, a homogeneous wave equa- tion. Evidently, the same wave equation is satisfied by the density p itself and the pressure p: 25 ce =O, 2B oY, Usually an additional function, known as the velocity potential , is introduced by means of v= —gradg. This is always possible if the initial distribution of velocities is irrotational. From av Lge a c grad 4, it follows that vee, 8 = WO,0) — of arad fr, 7) dr. Therefore, if curl v(x, 0) = 0, then v(r,0) = —grad y and* =~ 2 ft = we) = ead [¥ +e f sdr| grad ». ‘The velocity potential also satisfies the wave equation. Indeed, ee Substituting this equation into avr + div = 0, we obtain nah Example. Sound waves are generated at one end of a long straight rectangular tube of cross section a X &. We wish to investigate the patterns of these waves inside the tube for various frequencies. * This statement defines p as ¥ +c fy 8dr. 89 PROPAGATION OF SOUND. HELMHOLTZ EQUATION 323 For this purpose we seck a solution g(r, () which varies harmonically with time: ef, = We @ = real), where w is the frequency (presumably arbitrary). Note that the complex form is chosen for convenience. The real or imaginary part of pcan be used for a real sohution. It follows that ¥(r} must satisfy the PDE vet ky k= w/c, known as the Helmholtz equation. Physical requirements also demand that ¥(r) satisfy the following boundary condition at the walls of the tube: where d$/an = (grad ¢ - no) is the normal gradient of ¥, mg being the unit vector normal to the surface. This condition is due to the fact that, at the boundary, the air particles can move only in a direction tangential to the (rigid) walls of the tube, Remarks 1. The problem is essentially an eigenvalue problem in three dimensions since the Helmholtz equation can be rewritten in the form V¥- mM Om Kk) with undetermined». tis analogous to the equation for X(x) in the case of the vibrating string (p. 292). 2. There are no intidal conditions for the function g(r, 1) because we have specified the time dependence. We are only looking for suitable eigenfunctions in terms of which we can (hopefully) expand a particular physically realizable solution y(, #). 3, Experimental evidence has led us to assume that the time factor is e~*!, Actually, we could have invoked the method of separation of variables. Assume that 6,9) - KOTO. aT a) Veen, 6) Sa = eT Then Tackting equation (b) first, we see that the possibility that > 0 will lead to exponential functions which are not acceptabte on physical grounds: If the source of sound maintains a steady outpot, then the velocity potential g cannot increase or decrease steadily af all points r, because this would cause the intensity of the sound to increase (decrease) every- where and contradict the conservation of energy. The possibility that — 0 would allow only T(e) = const (same reasoning); this is of no physical interest, since 8 = 4/91 = 0. Consequently, the only possibility is \ < 0, leading to the form assumed in the beginning. 324 PARTIAL DIFFERENTIAL EQUATIONS 8.9 We solve the Helmholtz equation by separation of variables. Let the coordinate system be chosen as in Fig, 8.15, with the tube extending in the z-direction. Set HH) = XQYONZO; ; then ~ ax ay az a) Ger h Get oO Gea hz, i with the provision that 4; + Az + Ag = —k?. The con- dition ay/an = 0 at the walls leads to the statements aX) = ay = ° alo. | Vhs Figure 8.15 The condition on X implies that either 4, = 0 and X = const, or Ai < 0 (set Aa = —k3) and X = cos (mrx/a), where m is a positive integer. Similar conclusions are drawn for ¥. It is convenient, then, to introduce the following functions of x and y: XmnlX, Y) = eos — Gan = 0,1,2,3,-..), IX nay cos which are labeled by a pair of indices involving all possible combinations of the two nonnegative integers m and. In this, they are similar to the functions Yan of the preceding section; note, however, that the values m = 0 and n = 0 are now permissible. These functions are related (see below) to different modes of propagation of the sound atong the tube. ‘Turning now to the equation &Z an haZ, we see that the constant X, must satisfy 2 22 yee w we ar a= RP $b a) = -(4-"F-4). ‘Thus for a fixed frequency «, the constant 3 will be positive or negative depending on the selection of integers m and x, i.e., on the choice of the mode of propagation. If 2g is positive, Z wilt be proportional to e*¥, (If sound is generated at z = 0 and the tube extends along the positive z-axis, the increasing exponential solution is unphysical and must be discarded.) We say that the sound wave is attenuated. If, on the other hand, d3 is negative, then, setting Ag = — kJ, we have i Z = Acosk3z + Bsink3z. Note that kg = »/w?/c? ~ m2x2/a? — 1242/62, In this case we say that the sound wave is propagated, PROBLEMS: 325 Tt is seen that if the frequency w is fixed, only a certain (finite) number of modes will be propagated, namely, those modes for which m and n are sufliciently small to allow wfc? > m*n?/a? + n?7n?/b% The (0, 0)-mode is always propagated. lt represents a plane wave in the z {wis independent of x, y and is constant over a plane z = const at a given instant of time). Low frequencies can be propagated only by means of a few (lowest) modes, high frequencies allow a large number of modes, For a given mode (m, 1 = fixed), the frequency given by pn = we m2/a? FE 3/b™ is called the cutoff frequency. If & > mn, the mode is propagated; if & < enns it is attenuated. Remarks 1. In this problem we have investigated only the background for representation of an arbitrary acoustic wave propagating down the tube. It is now reasonable to con- jecture that such a wave can be decomposed into waves of all frequencies (by means of a Fourier spectrum) and, moreover, a given frequency component can be decomposed into a number of allowed modes. If the frequency w is considered as an eigenvalue, then it exhibits degeneracy reflected in the number of modes. 2. There is a possil problem. ility of further degeneracy (if a/b is rational), as in the membrane BIBLIOGRAPHY Carsaw, H.S., and J.C. Jarcen, Conduction of Heut in Solids. New York: Oxford University Press, 1959. Frankun, P., An Inteaduction to Fourier Methods and the Laplace Transformation, New York: Dover Publications, Inc., 1949. Morse, P. M., Vibration and Sound, New York: McGraw-Hill Book Co., 1948. Sommenrern, A., Partial Differential Equations in Physics. New York: Academic Press, Inc., 1949. PROBLEMS 1. A stretched string of fength L is “plucked” at the point and its initial shape is given by 326 PARTIAL DIFFERENTIAL EQUATIONS a |. A heat-conducting cylindrical rod of length L is thermall; |. Suppose that in the preceding problem the The initial distcibution of velocities is zero, that is, 3u/d/(x, 0) = 0. Show that the motion of the string can be represented by 2 « AL! 1. * Sy sin 4 sin cos BE WD = SEB Ay EE OT When a piano string of length Lis struck by the piano hammer near the point x = £, it is assumed that the initial distribution of velocities is given by vocos™= =!) (tor jx — &1 <4). 0 (ward): Assume no initial displacement and show that the motion of the sfring is given by volx) = vod <1 sin (nx$/L) cos (red /2L) 40d in BE we fin 1 (ad/iP sin 77 cos PEE. L A wx, ) T ulated over its lateral surface and its ends are kept at zero temperature. Ef the initial temperature of the rod is ue = const, show that the temperature x(x, £) at any later time must satisfy au ate 0,1) = AL) =O, o x77 ae” 4,1) = w(L,) = 0, — ulx,0) = ito. Show that exactly the same mathematical formulation is applicable to the problem of cooling an infinite slab of thickness Z, initially at temperature uo and with its surfaces maintained at zero temperature. Solve this problem by separating the variables and show that ee LNE MAX ee a TENE ain = Does this series converge rapidly for smnalll or for large values of 7 I temperature distribution is an arbitrary function of x: s(x, 0) = f(x). Show why the method of Fourier finite sine transforms can be used immediately. Use this technique to develop the formula b 2 potters mf in tHE uy = FT z= é sin J, Msn Tat. Attempt to solve Problems 3 and 4 by application of the Laplace transform with respect to # and avoiding the Fourier series altogether. Comment on the difficulties, If you cannot find a way out, take the special case f(x) = x in Problem 4 and show that x _ E sinh (xv/s/a) > PROBLEMS 27 Now apply the technique of Problem 15, Chapter 5, to complete the solution. Does the obtained series converge rapidly for small or large values of ¢? Solve Prob- lem 3 by this method. Suppose that the rod (or the sab) of Problem 4 is insulated atx = Oandx = L. Set up the new boundary conditions and establish a suitable set of eigenfunctions. Show that the solution is now a 1 . 2 oe, = if Ode + Ey corre eos B8) cos a. 7. Consider an infinite heat-conducting rod stretching in the x-direction (—c O if the initial temperature at ¢ = 0 is prescribed: u(x,0) = f(x). Formulate the problem (PDE, initiat condition, boundary conditions). Which method of solu- tion would you select and why? Using an appropriate convolution theorem, show that _{” pat - tet ge Me) = [se a 8. Consider the following two problems: a) Determine the steady-state (independent of time) temperature distribution in an infinite rod with the cross section shown in Fig, 8.17(a) and the lateral surfaces maintained at the temperature given. b) Determine the electrostatic potential in an infinite cyfinder split lengthwise into four parts and charged as shown (Fig. 8.176). t= tg const Insulation " ve b $ Insulation u=0 @ ga en 4 Insulation p= +V¥ const (b) Figure 8.17 State the mathematical formulation of both problems. Produce a rigorous proof (by symmetry considerations) that the solution of the second problem provides us automatically with the solution to the first one. Show, in the most efficient manner, that the solution of the second problem can be given by = Sto #0, 8) = wate, Ly sin nd. 328 PARTIAL DIFFERENTIAL EQUATIONS 9, Show that the problem discussed in Example 2, Section 8.5, can be solved by applying the Fourier sine transform with respect to x. (Consult, if necessary, the solutions of Problems 6 and 8 in Chapter 7.] 10, Consider the semi-infinite heat-conducting medium defined by the region x > 0 . and arbitrary y and z, Let it be initially at zero temperature and fet its surface x = 0 have a prescribed variation of temperature u(0, 1) = f(#) (for ¢ > 0), Use a method of your choice to show that 22 tats) ule, ) = f _—__. avr 10 E = TPR (#fint: The result of Problem 12, Chapter 5, may be helpful} 11. The principle of superposition (Section 8.2) can be used in a variety of situations not mentioned so far. Suppose that in Problem 3 above, only the end x = L is kept at zero temperature, while at the end x = 0 a prescribed heat current density q (p. 298) enters the rod, a) Show that the problem may be formulated by St) dr, ou du au St, = Sab 0) = Rie ae MEN Ho FT b= const, — fx, 8) b) Instead of imposing the initial condition u(x, 0) = 40, find a steady-state (time- independent) solution #1(x) of the above problem. ©) Show that the sum of «i(x)and the solution of Problem 3 yields the solution sought in (a). 12, Consider a very long (infinite) heat-conducting bar of rectangular cross section, such as the one shown in Fig. 8.18. a) Ifthe facex = @ is kept at 2 constant tempera- ture uo and the other three faces are kept at zero temperature, show that the steady-state solution can be expressed by Ty oN y, Isinh (axx/b) ney, u(x, y) = 5, nainh @era/b) 5 | b) Write (without solving the PDE) the steady. = [~~ state solution for the case in which the face » = bis kept at the temperature uo and the Figure 8.18 remaining three faces are al zero temperalure. ¢) Suppose that both faces x = aand y = bare kept at temperature wo, and the faces x = Oand y = Oat zero temperature. Can you write the solution immediately ? d) State clearly what features of the PDE, boundary conditions, etc., permit the use of the techniques hinted at above. i 13. For some additional applications of the superposition principle, consider Problem 12(c) again. a) Observe that the function w(x, y) = wox/a satisfies the PDE; state also the boundary conditions satisfied by 11. 14, . A stretched string of length L is initially at rest. PROBLEMS == - 329. b) Suppose the solution to Problem 12(c) is sought in the form wy y) = sex + uals, »). What boundary conditions should a(x, ») satisfy? ©) If u(x, ») is sought in the form of a Fourier sine series, show that ur, 3) = +2 x sinh (txy/a) + (—1)" sinh [real — Ma nsinh (rr6/a) d) Show that the function u(x, 9) given above satisfies ux, 0) = fx, 6) = (Hint: Develop the function uox/a in a Fourier sine series for 0 < x < a.) A radioactive gas is diffusing into the atmosphere from contaminated ground so that 1 pgm/cm? - sec escape into the air. Assume the ground and the atmosphere to be semi-infinite media with x = 0 as the boundary. a) Show that the density of the radioactive gas in the air is governed by the relations 90) pd en Wiew) — rel, (DA = const), a ox? = —X= const (pl, ) < »). b) Assume that p{x,0) = 0 (x > 0) and solve the problem by the Fourier cosine transform method to show that x _ (weg yt ply) = % f ep ote ak ¢) Solve the problem by the Laplace transform method and show that — 7 Dd Vo nye —2t/4Dy (x, f= X Bf —e @ dr, ot i? lo V5 [Hint: Obtain the inverse transform of (/V3)e and use the convolution theorem. } 6) Show that the solutions found under (b) and (¢) are equivalent, Hint: Recall that “ RR f ee g, Lae . ‘oO Dk2 +> from Problem 12, Chapter 5, vibrations described by the PDE Oe 1 oe St ae = he a8 3 oe 330 PARTIAL DIFFERENTIAL EQUATIONS 17, 1 Find the motion of the string by the following method of complementary function: Seek a solution v = ui + w2, where a1 is a general solution of the homogeneous PDE (as above, with the right-hand side of the equation being zero) and satisfies the boundary conditions but not the initial conditions, while u2 is a particular solu- tion of the nonhomogencous PDE and the boundary conditions but is independent of time. Complete this idea; show its consistency and solve the problem. A rectangular membrane 0 < x < a, 0 < y < 6 is clamped on all sides and is loaded by a uniformly distributed external force q (per unit area). 2) Show that the static (time-independent) deflection u(x, ») satisfies a nonhomoge- neous PDE and homogeneous boundary conditions. b) Show that the function qx(@a — x) ms y) = “ar satisfies the PDE, but not all the boundary conditions, } If the overall solution is sought in the form a1(x, »} + walx, ») (see Problem 15), determine the PDE and the boundary conditions which should be satisfied by i2. d) Complete the solution, simplify it, and show that 2 axle — 9) _ fact S.A cosh inetay — b9/2a) mex Wn) = SF ys cosh (ax6/2a) a ‘The stretched string of length L undergoes forced vibrations according to au _ oe % u Sa TO gga t Asner (A = const). a) Seek a steady-state solution in the form u(x, £) = y(x)sin wt and show that it can be written as sn 1) me 24 Sin sin EL — 29/2) in a (=<): ® cos (KE/2) b) Let the string be at rest prior to ¢ = 0. Solve tie problem by the finite transform method and show that ule, = 24 sin ot x sin(nex/L) _ 4A Lan sin (ors/L) in ot, Gy x eda neo — os 5 OR nod (ww, — wo) where w, = mxrc/L; assume that w cy. Comment on the physical meaning of each sum. ¢) Shovid the first sum in (b) be identical with the solution () in (a)? Verify your answer by expansion of (i) into a Fourier sine series. d) Suppose that « = nixc/L for some m. What modifications are needed in part (0)? Solve this problern and comment on the validity of the solution. : . Consider a parallelepiped a X & X ¢ made of solid material. One of the (a X 6)- faces is maintained at the temperature # = uo = const, and the other five faces are kept st zero temperature. Suppose that we wish to obtain a time-independent tempera- ture distribution, 19. OBLEMS = 331. a) Formulate the problem in cartesian coordinates with one vertex of the paral- Telepiped at the origin. b) Solve the problem and show that your solution is equivalent to (or identical with) 169 = <> 1 sinhams(c — 2). mex, nny uly y 2) = = me sin — sin == Ang gg tat Sih ermne a where am, = 9 mi2/a2 + 12/b2, ©) Evaluate a(x, y, 2) = lime. ux. y, 2). d) The function ax, ¥, 2) can be used to solve a certain problem ia electrostatics. Formulate this problem. A rectangular membrane with dimensions (a X 6) is at rest for ¢ <0. For 7 > 0, it is subject to external forces (per unit area) FO, ys) = Ax, ye a) Formulate the problem. b) Solve the problem for the general (no resonance) case ifw =f rev/m2Ja2 + n2/b2 with m, = integers. ©} Solve the problem for the case of resonance: = xeV/mg/a? + 1§/6? (for some particular mp and sto). & Comment on the validity of the solution found in (c). e) Discuss the resonance case if the frequency is degenerate, ). We wish to investigate the characteristic modes of vibration of air in an enclosure of the shape of a rectangular parallelepiped (~a/2 < x < a/2, —b/2 < y < b/2, -e/2 S21 < c/D, a) Set up the PDE and the boundary conditions satisfied by the velocity potential, b) Secking solutions varying in time as et, determine the characteristic frequencies and the characteristic modes of vibration. c) Leta — 6 = mand c = 3m. Show that the lowest triple-degenerate mode has frequency » — wev/2 (c in m/sec). What are the lowest fourfold and fivefold degenerate modes?

You might also like