You are on page 1of 485
a a introduction to modern electromagnetics pose eS ee Carl H. Durney Associate Professor of Electrical Engineering University of Utah Curtis C. Johnson Associate Professor of Electrical Engineering University of Washington ROBERT E. KRIEGER PUBLISHING COMPANY MALABAR, FLORIDA 1982 Original Edition 1969 Reprint Edition 1982 Printed arid Published by ROBERT E. KRIEGER PUBLISHING COMPANY, INC. KRIEGER DRIVE MALABAR, FLORIDA 32950 Copyright © 1969 by McGraw-Hill, Ine Reprinted by Arrangement All rights reserved. No part of this book may be reproduced in any form or by any electronic or mechanical means including information storage and retrieval systems without permission in writing from the publisher. Printed in the United States of America Library of Congress Cataloging in Publication Data Durney, Carl H., 1931- Introduction to modern electromagnetics. Reprint. Originally published: New York : McGraw- Hill, 1969. Bibliography: p. Includes index. 1. Electromagnetic fields. 2. Boundary value problems. 3. Electric engineering. 1. Johnson, Curtis C. IL. Title. QC665.E4D87 1982 530.1°41 81-23602 ISBN 0-89874-333-8 AACR2 To Marie To Katharine and Wilma preface Electromagnetics continues to be one of the most fundamental subjects in an engineering curriculum because it provides a basic physical and mathe- matical understanding of electric, magnetic, and propagation phenomena. The principles of electromagnetics have been applied in an increasing number of research areas in recent years, placing an emphasis on sound- ness, breadth, and the dynamic aspects of an undergraduate course in electromagnetic field theory. For example, the microwave interest with its base in devices, antennas, and propagation, has moved into microwave phenomena in solid-state materials and gaseous plasmas, where statistical and quantum-mechanical effects are prominent. Interest in the optical spectrum has grown rapidly in step with laser advances into new material interactions, acoustic interactions, propagation, and nonlinear effects. These new research and development interests require a concise and modern undergraduate introduction to electromagnetics. A survey was conducted recently to determine the present-day needs for an electromagnetics textbook. The results indicated a need for mathematical and physical clarity, the use of examples and applications, and a problem-solutions manual to reduce the time required by the instructor. This textbook was prepared with these objectives in mind. vii viii preface The book is designed for a first course in engineering electro- magnetics at the junior or senior level. There is enough material for a full year’s course, and some of the more difficult sections, which have been marked with a star at their beginnings and endings, are optional and can be omitted with no loss of continuity. This makes the text some- what adaptable to courses of varying difficulty and length. Problems at the end of each chapter are numbered according to the section to which they are pertinent. Thus Problem 1.5-1 is the first problem of Section 1.5. A number of problems are designed to be solved by computer. We have assumed that students using this book have been intro- duced to vector analysis. Although an extensive review of vector analysis is given in Chapter 1 and the text is mathematically self-con- tained, the reader who has not had some previous experience with vector analysis will have difficulty learning the mathematics fast enough. Chapter 1 is meant to be a complete but compact discussion of vector analysis. Since many students have had little experience with cylindri- cal and spherical coordinate systems, these are discussed in some detail. Physical interpretations of operations such as the divergence and curl are emphasized. We have tried to make the mathematical level high enough to prepare the student for graduate courses in electromagnetism, while at the same time emphasizing concepts and physical understanding. Our approach differs principally from some of the more traditional approaches in that dynamics is emphasized over statics. Since field con- cepts are more easily introduced in terms of statics, the basic electrostatic and magnetostatic concepts are developed first in Chapter 2. The extension to time-varying fields is made, and Maxwell’s equations are developed in Chapter 3. The remainder of the text is devoted to under- standing, solving, and applying Maxwell’s equations. The rationalized mks system of units is used throughout the text. In Chapter 4 energy and power concepts are discussed. This is fol- lowed in Chapter 5 by a description of the interaction of electromagnetic fields with charges in materials. The purpose of Chapter 5 is to show in a detailed way how the interaction between the charges in materials and electromagnetic fields can be described in terms of permittivity and permeability. Since the interaction is easier to understand in the case of dielectrics and conductors than magnetic materials, dielectrics and conductors are emphasized; they are treated from a common base and in terms of time-varying fields. Chapter 6 contains a discussion of the wave equations, potentials, and planewaves. This chapter presents the fundamental concepts of wave propagation. Refraction and reflection at dielectric boundaries, reflection at conducting boundaries, skin effect, radiation from antennas, Preface ix and propagation between parallel conducting plates are some of the topics considered. Some examples involving laser beams are given. The connection between field theory and circuit theory is discussed in Chapter 7. The quasi-static approximation is used to show how Kirchhoff’s laws are related to Maxwell’s equations. This helps to show the connection between planewave propagation, described in Chapter 6, and the wave propagation on two-conductor systems, described in Chapter 8. Transmission-line concepts such as impedance, reflection coefficient, and standing-wave ratio are discussed in Chapter 8. Also included are the Smith chart and optional sections on tapered trans- mission lines and transients on transmission lines. The last chapter is on boundary-value problems. In this chapter mathematical techniques for solving the wave equations are discussed. Propagation in wave guides is treated as a boundary-value problem. An optional section describes how a partial separation of variables in Max- well’s equations leads to a description of wave-guide propagation in terms of transmission-line equatidns. Then wave-guide propagation is dis- cussed in terms of transmission-line concepts, including the use of the Smith chart. The development of this text was influenced by many sources. We are grateful to the students who worked through the text when it was in the form of course notes and offered corrections and valuable suggestions. Our colleague, Professor Om P. Gandhi, and several reviewers have also offered many valuable suggestions. We appreciate the excellent work of Mrs. Marian Swenson and her staff in typing and reproducing the. manuscript through several revisions. Carl H. Durney Curtis C. Johnson contents CHAPTER 1 11 1.2 13 14 15 MATHEMATICAL INTRODUCTION Introduction 1 Review of Basic Properties of Vectors 2 Vector Addition and Subtraction 3 Vector Products 4 Unit Vectors 5 Scalar and Vector Fields 5 Coordinate Systems 7 Rectangular Coordinate System 8 Cylindrical Coordinate System Bet Spherical Coordinate System 13 Addition and Multiplication of Vectors Represented in the Three Standard Coordinate Systems 15 Generalized Orthogonal Curvilinear Coordinates 19 xii contents i 18 19 1.10 111 WL 1.18 1.14 1.15 1.16 1.17 1.18 CHAPTER 2 2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 CHAPTER 3. 3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 Transformations between Coordinate Systems 19 Rectangular-to-Cylindrical Transformation 19 Rectangular-to-Spherical Transformation 23 Transformations from Another Point of View 24 Differentiation of Vectors 25 Integration of Vectors 29 Line Integrals 29 Surface Integrals 35 Volume Integrals 39 Gradient 39 Divergence 41 Curl 45 Derivation of an Expression for the Curl of a Vector 48 The Divergence Theorem (Gauss’s Theorem) 53 Stokes’s Theorem 55 Other Operations Involving V 56 Complex Vector Notation 59 Source-point and Field-point Notation 61 The Delta Function 64 ELECTROMAGNETIC FIELDS Introduction 71 The Electric Field 73 Coulomb’s Law 76 Electric Potential 84 ‘The Magnetic Field 87 Current Relations 89 Ampére’s Law of Force 95 Magnetic Field-generating Systems 99 The Magnetic Dipole 101 Review of the Force and Field Relations 105 ELECTROMAGNETIC FIELD LAWS Introduction 108 Gauss’s Law 109 Applications of Gauss’s Law i The Magnetic Source Law 114 Faraday’s Law 115 A Stationary Loop in a Time-varying Magnetic Field 7 A Moving Loop in a Constant Magnetic Field 119 A Moving Loop in a Time-varying Magnetic Field 122 3.9 3.10 3.11 3.12 3.13 CHAPTER 4 4.1 4.2 4.3 44 4.5, *4.6 47 4.8 CHAPTER 5 5.1 5.2 5.3 5.4 5.5 5.6 5.8 5.9 5.10 5.11 5.12 5.13 contents xiii The Continuity Equation 125 Ampére’s Circuital Law 126 Derivation of Ampére’s Circuital Law from the Biot-Savart Law 129 Displacement Current and the General Time-varying Form of Ampére’s Circuital Law 131 Displacement Current 133 Maxwell’s Equations 135 PHYSICAL PROPERTIES OF FIELDS Introduction 140 Field Force on Charges 141 Energy Associated with an Assembly of Electric Charges 148 Power-Energy Conservation Theories 149 Complex Power and Energy 153 The Electric Force Field 156 Derivation of the Electric Surface-force-Density Relation 162 The Magnetic Force Field 163 Derivation of the Magnetic Surface-force-Density Relation 166 ELECTROMAGNETIC FIELDS AND MATERIALS Introduction 171 Interaction of Charges and Fields 172 Effects of Fields on Nonmagnetic Materials 175 Polarization of Bound Charges 175 Drift of Conduction Charges 182 Effects of Nonmagnetic Materials on Fields 183 Polarization Charge Density and Current 183 Conduction Current 190 The Displacement Vector D and Permittivity « 190 Dielectrics 194 Conductors 196 Semiconductors 201 Complex Permittivity for Sinusoidal Time Variation 201 Magnetic Materials 203 Magnetization-current Density 204 Magnetic Field Intensity H and Permeability 208 Magnetic Scalar Potential 210 Materials and Maxwell’s Equations 213 Boundary Conditions 214 Normal Components 214 Tangential Components 216 xiv contents 5.14 5.15 CHAPTER 6 6.6 6.7 *6.8 6.10 6.11 6.12 6.14 6.15 6.16 6.17 CHAPTER 7 Ee 7.2 Interdependence of Boundary Conditions of Normal and Tangential Components 217 Boundary Conditions for Good Dielectrics 218 Boundary Conditions for Perfect Conductors 218 Boundary Conditions for Magnetic Materials. 219 Summary of Boundary Conditions 219 Materials and Power and Energy 222 Power Dissipation in a Dielectric with Losses 224 THE WAVE EQUATION AND PLANEWAVE PROPAGATION Introduction 232 One-dimensional Electromagnetic Planewaves 233 Inhomogeneous Wave Equations for the Fields 238 Potential Functions 240 Potential-function Wave Equations 242 The Lorentz Condition 244 Retarded Potentials 244 Retarded Potentials as Solutions to the Potential Wave Equations 245 The Electric Dipole Antenna 247 The Two-dipole Array 250 Power Radiation from the Electric Dipole Antenna 253 General Planewave Solution to the Field Homogeneous Wave Equations 256 Field Relations in a Planewave 258 Power and Energy in a Planewave 259 Planewave Reflection at a Perfect Conductor: Normal Incidence 261 Planewave Reflection at a Perfect Conductor: Oblique Incidence 264 Planewave Incidence upon a Dielectric Interface 266 Wave Propagation in Homogeneous Dielectrics 272 Effects of Conduction Current in Dielectrics 274 Wave Propagation in Conductors 275 Tenuous Gaseous Conductors 276 Solid-state Conductors 276 The Skin Effect 281 CIRCUIT THEORY FROM FIELD THEORY Introduction 286 Quasi-static Fields 287 7.3 74 75 7.6 CHAPTER 8 8.1 8.2 8.3 8.4 8.5 8.6 8.7 8.8 8.9 8.10 8.11 8.12 contents xv Fundamental Circuit Relations 288 The RL Circuit 288 The RLC Circuit 290 A Generator and an External RL Circuit 292 Lumped-circuit Theory 295 Kirchhof’s Current Law 296 Inductance 297 Mutual Inductance 298 Inductance and Energy 300 Finite Conductors 303 Internal Impedance 305 Capacitance 308 Capacitance and Energy 313 TRANSMISSION OF ELECTROMAGNETIC ENERGY Introduction 316 Diffraction of Planewaves 318 The Distributed-circuit-element Transmission Line 322 The Lumped-circuit-element Transmission Line 327 Normal-mode Analysis of the Transmission-line Equations 328 Transmission-line Progagation Characteristics 332 Characteristic Impedance 333 Propagation Constant 335 Lossless-transmission-line Discontinuities and Impedances 339 Lossy-transmission-line Discontinuities and Impedances 349 The Smith Chart 351 Lossless Lines 352 Lossy Lines 360 Impedance Transformation and Matching 360 Reactive Elements 361 Matching with Shunt Elements 362 Matching with Series Elements 363 Quarter-wave Transformers 364 Half-wave Transformers 365 Nonuniform and Tapered Transmission Lines 365 Low-reflection Tapered Transmission Lines 368 Exponential Impedance Taper for Matching 368 Transmission-line Transients 371 The Laplace Transform of the Transmission-line Equations 371 Quiescent Lines 378 xvi contents CHAPTER 9 9.1 9.2 9.3 94 9.5 9.6 9.7 9.8 9.9 9.10 *9.11 9.12 *9.13 APPENDIX bibliography index 461 BOUNDARY-VALUE PROBLEMS Introduction 386 A One-dimensional Boundary-Value Problem 387 Fields between Infinite Parallel Plates 389 Separation of Variables in Rectangular Coordinates 394 Orthogonality and Eigenfunction Expansions 397 Uniqueness of Solutions for Field Equations 400 Method of Images 402 The Parallel-plate Wave Guide 405 Solution of Maxwell’s Equations 405 Satisfying the Boundary Conditions 409 Parallel-plate-wave-guide Modes 410 Propagation Characteristics of the Modes 411 Field Configuration of the Modes 412 Separation of Maxwell’s Equations 414 The Rectangular Wave Guide 415 Derivation of Transmission-line Equations for Wave Guides 420 Transmission-line Equations for TEM Modes on Guiding Structures 422 Transmission-line Equations for TEM Modes in Free Space 427 Transmission-line Equations for TE and TM Modes 428 Transverse Boundary Conditions on H, and E, 433 Transmission-line Characteristics of Wave Guides 437 Characteristic Impedance and Propagation Characteristics 438 Wave-guide Transmission-line Discontinuities and Impedances 440 Transmission-line Power and Energy 445 Energy-flow Velocity (Group Velocity) 450 The k8 Diagram 451 VECTOR RELATIONS 455 459 introduction to modern electromagnetics CHAPTER 7 mathematical introduction 1.1 INTRODUCTION Vector calculus is not strictly necessary to the study of electromagnetic field theory, but without it the presentation of the principles of electro- magnetic theory would be awkward and cumbersome. Vector calculus not only provides compact and meaningful expression, it also provides a powerful means of comprehension. Hence the use of vector calculus in field theory has become standard. It is assumed that the reader has previously obtained a basic under- standing of vector analysis. Consequently, this chapter is not meant to be a rigorous mathematical development of the principles of vector calcu- lus. Rather, it is meant to provide a review of important principles which are used repeatedly throughout the text, and to provide a common basis and point of departure for the study of the remainder of the text. Hence application and interpretation rather than rigorous mathematical proofs are stressed. Some of the very elementary principles are included for convenience of the reader and for completeness. Since it is difficult to provide a good physical interpretation of the mathematical relations before the corresponding physical principles are 1 2 introduction to modern electromagnetics chapter one introduced, the reader may not feel at first that he has gained a good appreciation for all the points developed in this chapter. It is hoped, however, that this chapter will provide a good foundation and that a deeper understanding and appreciation will be obtained as the relevant physical principles are encountered. 1.2 REVIEW OF BASIC PROPERTIES OF VECTORS In contrast to a scalar, which is characterized by only a magnitude, a vector is characterized by both a magnitude and a direction. A familiar example of a vector quantity is velocity. It is convenient to represent a vector quantity graphically as a directed line segment, as shown in Fig. 1.1. The length of the line segment represents the magnitude of the vector, and the arrow represents the direction of the vector. It is sometimes also convenient to think of the displacement of a point as representing a vector quantity. The magnitude and direction of the displacement then represent the magnitude and direction of the vector. In this book vectors are denoted by boldface type, as indicated in Fig. 1.1. The basic vector operations, which are only briefly reviewed here, are addition, subtraction, and two kinds of multiplication—the dot and cross products. It is important to note that these operations are inde- pendent of any coordinate system which might be used to locate the vec- tors in space. A detailed discussion of the representation of vectors in several coordinate systems is given in Sec. 1.4, but the basic operations are defined independently of any coordinate system. Before discussing the basic operations, we need the definition of “equal” vectors. Any two vectors which have the same directions and the same magnitudes are equal vectors. Because of this definition, we are free to “slide” a vector to any position in space if we do not change its direction or magnitude. For example, the vector A in Fig. 1.1 may be used to denote the velocity of a projectile. Although it is customary to locate the base of the vector at the projectile, it is not necessary to do so. The length and direction of the vector specify the velocity, not the loca- tion of the vector in space. Thus any of the vectors A shown in Fig. 1.2 may be used to denote the projectile velocity. In some instances, the idea of being able to slide a vector is confusing. For example, suppose the wind velocity in some region of space is to be represented by vectors. This could be accomplished by assigning a vector to each point in space Fig. 11 A graphical representation of a vector: a directed line segment. . sec. 12 mathematical introduction 3 Projectile Fig. 12 Any of the vectors A shown A may be used to denote the projectile velocity. (see Sec. 1.3). The vector at a given point would then represent the wind velocity at that point. In this case it would be confusing to think of sliding the vector from one point to another because it is important to associate one certain vector with a given point. In cases like these we do not usually think of sliding vectors, although we could if we were careful somehow to keep track of which vector belongs to which point. VECTOR ADDITION AND SUBTRACTION The addition of two vectors may be conveniently defined in terms of the displacement of a point. Thus, if A represents the displacement from point 1 to point 2 and B the displacement from point 2 to point 3, as shown in Fig. 1.3, then C represents the displacement from point 1 to point 3. It is obvious that C is equivalent to A + B; that is, A+ B=C. The sum of two vectors may thus be found by a graphic construction. It is easy to show. by such graphical construction that A+B=B+A (1.1) and A+(B+C)=(A+B)+C (1.2) These two characteristics of vector addition are called the commutative law and the associative law of vector addition, respectively. The negative of a vector is defined as a vector with the-same mag- nitude but opposite direction. With this definition, vector subtraction is easily defined in terms of vector addition, A-B=A+ (-B) (1.3) 3 Fig. 1.3 The sum of two vectors, A 4 Introduction to modern electromagnetics chapter one VECTOR PRODUCTS As constrasted with a single kind of scalar mutilplication, two kinds of vector multiplication are defined. They are the dot and cross products, also called the scalar and vector products of two vectors. The dot product of two vectors is a scalar quantity and is defined as A-B= ABcos6 (1.4) where 6 is the smaller angle between the two vectors, as indicated in Fig. 1.4, The notation A means the magnitude of the vector A. The dot product is commutative and distributive; that is, A-B=B-A (1.5) A-(B+C)=A-B+A-C (1.6) The cross product of two vectors, written A x B, is a vector whose magnitude is AB sin @ and whose direction is perpendicular to the plane defined by A and B and is in the direction a right-hand screw would travel when turned in the same direction as A turned into B through the angle 6. The graphical representation of AxB=C is shown in Fig. 1.5. It is easy to show from the definition that the cross product is not commutative, but that it is distributive; that is, AxB#BxA but Ax(B+C)=AxB+AxcC Some vector identities involving dot and cross products of more than two vectors are listed in the Appendix. B 8 A Plane of A and B Fig.14 The angle be- Fig. 1.5 The vector cross product A X B tween two vectors = C, where C is perpendicular to A and B. used in the dot prod- uet. sec. 13 mathematical introduction 5 UNIT VECTORS A unit vector is defined as a vector of unit magnitude. We can easily obtain a unit vector in the direction of any given vector by dividing that vector by its magnitude. Thus the unit vector A in the direction of A is given by A Br In general, unit vectors will be designated in this book by carets. We can also write any vector in terms of its magnitude and a unit vector, A=AA It will often be convenient to represent vectors in this manner; this idea is extended in Sec. 1.4, where vectors are written in terms of three unit vectors called base vectors. These base vectors are useful because of their convenient relations to coordinate systems. It is also convenient, as dis- cussed in Sec. 1.4, to write addition, subtraction, and dot and cross prod- ucts in terms of these base vectors. As a final definition in this section, let us consider the multiplication of a vector by a scalar. The product of a vector and a scalar is defined as a vector whose direction is the same as that of the vector being multi- plied and whose magnitude is the product of the scalar and the magnitude of the vector; that is, if A= AA then A = OAA (1.7) 1.3 SCALAR AND VECTOR FIELDS Since the concepts of vector and scalar fields are used frequently in this book, a careful discussion of these concepts is given in this section. Let us first review the definition of a scalar point function. Consider a given set of points, such as all the points in space. Suppose there is a certain scalar value corresponding to each point. The class of points is called the domain of definition, and the correspondence of scalar values to each point is called a scalar point function. The scalar values corresponding to the points are said to be the range of values of the scalar point function. A scalar point function is written as f(P), where P stands for any point in the domain of definition and f is the scalar value at the point P. An example of a scalar point function is the temperature at each point in space. A scalar point function is often called a scalar field. An often-used graphical representation of a scalar field is one in which the equal values of the scalar point function are connected by lines, as 6 introduction to modern electromagnetics chapter one 100 Fig. 1.6 Graphical representation of a scalar field. The lines are constant values of the scalar point function. shown in Fig. 1.6. In the representation of a temperature field, the lines of constant temperature are called isotherms. Geographical contour maps are representations of the scalar point function whose range of values con- sist of the distances above sea level. A scalar field we shall be often con- cerned with in this book is the potential field. The lines of constant potential are called equipotentials. A vector point function is a point function whose range of values is a class of vectors. Thus to each point in the domain of definition there corresponds a vector. We write a vector point function as F(P), where P stands for any point in the domain of definition, and F is the vector at the point P. As you have probably already guessed, a vector point func- tion is often called a vector field. Wind velocity is an example of a vector field. Ifa fan generates a continuous flow of air in a room, the magnitude and direction of the wind at each point P may be determined. The resultant vector wind velocity v(P) is a vector field function describing the magnitude and direction of the wind velocity at every point in the room. The vector fields we shall be most concerned with in this book are the electric and magnetic fields. The graphical representation of vector fields is not as easy as that of scalar fields. One representation of a vector field is shown in Fig. 1.7. The vectors represent the magnitude and direction of the vector field at ee eae | Fig. 1.7 One method of representing a vector field. sec. 1.4 mathematical introduction 7 a Fig. 1.8 Fluid flow around a cylinder. This illus- trates another, probably better, method of graphi- cally representing vector fields. points P located at the base of each vector. As noted in Sec. 1.2, we do not usually think of “sliding” a vector in connection with vector fields, because it is important to maintain the correlation between each point and the one vector which has been assigned to that point. Hence, when a vector is part of a vector field, we usually think of it as “bound” to a given point. It turns out that the kind of representation shown in Fig. 1.7 is not as convenient as might be expected, because the symmetry of the field is often obscured and the field is indicated only at discrete points. A more commonly used method is one in which continuous lines are drawn, with arrowheads indicating a direction. The direction of the field at each point is then tangent to the line, which is called a flow line, or a fluc line. The magnitude of the field is indicated by the separation between lines. If the lines are far apart, the vector magnitude is small. An example of this is the velocity field of fluid flow around a cylinder, as illustrated in Fig. 1.8. Although many kinds of vector fields represent the flow of particles, such as fluid molecules or charged particles, there are many vector fields which do not represent an actual physical flow of particles. It is often convenient, however, to think of these fields as representing fluid flow, even though there is no physical motion involved. The field lines are therefore quite generally called flux lines, and the field is spoken of as being a flux density. The total flux (flow) through 'a surface would then be found by an integration procedure. This kind of integration is dis- cussed in Sec. 1.7. 1.4 COORDINATE SYSTEMS As was pointed out in Sec. 1.2, vectors and the basic vector operations are defined independently of any coordinate system. It becomes very con- venient, however, to use coordinate systems in solving practical problems and for the understanding of physical principles. A coordinate system serves two purposes: one is to locate points in 8 introduction to modern electromagnetics chapter one space, and the other is to furnish a convenient way to represent the magni- tudes and directions of vectors at points in space, that is, to represent vector fields. This is done in terms of a set of three unit vectors which are called base vectors. The expression for a vector field in terms of these base vectors of a coordinate system is called a representation of the vector field. For convenience, the base vectors are usually chosen to be mutually orthogonal (mutually perpendicular), but it is not necessary that they be so. We shall first discuss in detail the three most commonly used coor- dinate systems—rectangular, cylindrical, and spherical—and then discuss the general orthogonal curvilinear coordinate system. The choice of coordinate system depends upon the symmetry of the problem at hand. For example, in cases where cylindrical symmetry is present, the use of a cylindrical coordinate system rather than a rectangular system greatly simplifies the mathematics. RECTANGULAR COORDINATE SYSTEM The rectangular coordinate system is the most familiar of the three stand- ard coordinate systems. It should be used whenever possible because of its simplicity and, as we shall see later, because the base unit vectors are constants. Figure 1.9 shows the three mutually orthogonal axes, which we have labeled the x, y, and z axes. Also shown are the three base vec- tors, labeled %, §, and 2, which lie along the z, y, and z axes, respectively. Since these base vectors have constant magnitude and direction, they are constant vectors. In some other coordinate systems, some of the base vectors are not constant because their directions are variable. A right-handed system is defined as one in which x § = 2. Any point in space is located by its values of x, y, and z. For example, point 1 in Fig. 1.9 is located at x = a, y = yi, and z = 2, as indicated. We say that x, y, and z are the independent variables, and any function, such as a scalar or vector point function, which in general depends on 2, y, and z (is a function of x, y, and z) is called a dependent variable. Fig. 1.9 Rectangular coordinate system. sec. 1.4 mathematical introduction 9 Fig. 110 Rectangular components of A. In general, vector or scalar point functions vary from point to point in space. Since each point in space can be located by its z, y, and z coor- dinate, a vector or scalar point function can be written as a function of z,y,andz. Itis easy to see how a scalar field can be expressed in terms of z,y, and z. For example, the temperature scalar point function T(P) is simply T(z,y,z) when (z,y,z) corresponds to the point P. A vector field is more complicated, since both the magnitude and direction may be func- tions of z,y,andz. However, a vector field can be conveniently expressed in terms of its components along the base vectors. If a known vector field A is a function of x, y, and z, we may write A(z,y,2) = Az(x,y,z)& + Ay(x,y,2)9 + A.(x,y,2)2 (1.8) where A,(z,y,z), A,(z,y,2), and A,(z,y,z), which are the scalar components of A along the z, y, and z axes, respectively, are to be determined. Equa- tion (1.8) is called the representation of the vector A in a rectangular coor- dinate system. Since the base vectors are constants, the dependence of both magnitude and direction of A on 2, y, and z is completely contained in the dependence of A., A,, and A, on z, y, and z. Expressing A by Eq. (1.8) is commonly called “writing a vector as a sum of its components along the base vectors.” A geometric representation based on the addi- tion of vectors is shown in Fig. 1.10. Since the direction and magnitude of A are known, A,, A,, and A, can be obtained by scalar multiplication of Eq. (1.8) in turn by &, 9, and 2. Because %, §, and 2 are mutually orthogonal, the results are &-A = A,(z,y,2) (1.9) F-A = A,(x,y,2) (1.10) 2-A = A,(z,y,z) (1.11) The advantage of having orthogonal base vectors is obvious here. 10 introduction to modern electromag: chapter one y x Fig. 1.11 A vector in the zy coordinate system. Example 1.1 Finding the components of a vector A nearly trivial example which illustrates these principles is illustrated in Fig. 1.11. A known vector A lies in the zy plane at an angle @ from the z axis. Using (Eq. 1.4), we immediately find that A, =%-A = A cos 0 Ay=9-A © = A coe (F- ) = Asin 6 Here A, and A, are constants, but in general they may be functions of z, y, and z. A less formal way of obtaining A, and A, is simply to resolve A into com- ponents along the z and y directions by finding two vectors in the direction of and ¥ which add up to A. The two methods are equivalent. In connection with integration, which is discussed in detail in Sec. 1.5, we need the differential arc length expressed in rectangular coor- dinates. A differential length results when one of the coordinates is increased by a differential amount. In rectangular coordinates the differ- ential lengths are dz, dy, and dz. The total differential arc length dl is the result of increasing each coordinate by a differential amount. It is given by (dl)? = (dz)? + (dy)? + (dz), where dl corresponds to the length of the diagonal of the box shown in Fig. 1.12. Another quantity of interest is the differential volume, dV = dzdy dz, which is illustrated in Fig. 1.12. Fig. 1.12 Differential lengths and volume in rectangular coordinates. sec. 1.4 mathematical introduction 11 As a prelude to the definitions of the unit vectors in the cylindrical and spherical coordinate systems, note that a more general definition of %, §, and 2 can be made in terms of the surfaces of constant coordinate value. Thus & is defined as a unit vector perpendicular to the surface x = constant and in the direction of increasing x. Similar definitions are made for § and 2. It should also be noted that the point (@1,y1,21) is defined as the intersection of the three mutually orthogonal surfaces =f, y =y, andz =z. CYLINDRICAL COORDINATE SYSTEM The independent variables of the cylindrical coordinate system are r, g, and z. For convenience, the cylindrical coordinate system is shown in Fig. 1.13 with respect to a rectangular coordinate system. The cylindrical coordinate system is used in place of the standard rectangular coordinate system for problems with cylindrical symmetry. When cylindrical struc- tures, such as a round pipe or rod, are analyzed, the mathematics is much simpler in cylindrical coordinates. Ther = constant surface is a cylinder concentric with the z axis, the y = constant surface is a plane passing through the z axis, and the z = constant surface is a plane perpendicular to the z axis. The intersection of these three surfaces defines the point (r,¢,2). The unit base vectors are f, §, and 2. Each base vector is defined at a coordinate point as being perpendicular to the constant coordinate surface and in the direction of increasing coordinate. These vectors are shown in Fig. 1.14. Since the three constant coordinate surfaces are mutually orthogonal, it follows from the definitions that the three base vectors are mutually orthogonal. A right-handed coordinate system is ‘Constant z z em ~~, ot one Constant r cylinder Fig. 1.13 The cylindrical coordinate system. 12 introduction to modern electromagnetics chapter one Fig. 1.14 The base vectors of the cylindrical coordinate system at the coordinate point x (rez). defined by # x § = 2. Note that and § are always parallel to the zy plane. A very important characteristic of f and 4 is that their directions are functions of ¢ (their magnitudes, of course, are constant). This is in contrast to %, §, and Z, which are constant vectors. It is essential that this ¢ dependence of ¢ and § be kept in mind, especially in integrating and differentiating vectors. The fact that ¢ and § vary with ¢ is a nuisance which may be eliminated, of course, by using rectangular coordinates. Vectors are expressed in the cylindrical coordinate system in a way similar to that used for a rectangular system. The representation of A in a cylindrical coordinate system is A(r,9,2) = Ad(ry2)t + Ag(r,,2)8 +. Aa(r,e,2)2 The differential lengths are dr, r dy, and dz. Since ¢ is an angle, the differential length obtained by increasing ¢ by dg is given by r dg, as illustrated in Fig. 1.15. The total differential are length is (a)? = (ar) + (r dg)? + (de)? The differential volume, shown in Fig. 1.16, has the shape of a truncated wedge and is given by aV =rdrdg dz (1.12) because the truncated wedge approaches a rectangular parallelpiped as dr, r dg, and dz become very small. y 45 The differential length r ag in a cylindrical 4 nae coordinate system. sec. 1.4 mathematical introduction 13 dz Fig. 1.16 The differential volume 4 lp in cylindrical coordinates. x Le SPHERICAL COORDINATE SYSTEM In the spherical coordinate system, the independent variables are 0: and y. Unfortunately, it has become common practice to use the same symbol r for both the cylindrical and spherical coordinate systems, even though the variable is not the same in the twosystems. In most practical situations, however, there is little doubt as to the intended meaning of the symbol r. Since the reader will be faced with this problem in other lit- erature, it seems useless to try to improve the notation, and so we shall follow the common practice. The spherical coordinate system is used for problems with spherical symmetry and in radiation problems, where its use greatly simplifies the mathematics. The spherical coordinate system is shown in Fig. 1.17, with the rec- tangular coordinate system asareference. LetO represent the origin and P the point (r,6,9). Then r is the length of the line OP. The angle 6 is the angle between the z axis and OP, and the angle ¢ is the angle between the projection of OP on the zy plane and the z axis. Constant @ ae Constant r sphere Fig. 1.17 The spherical coordinate system. 14 introduction to modern electromagnetics chapter one Fig. 1.18 Base vectors of the spherical coordi- nate system. The r = constant surface is a sphere centered at the origin, the 6 = constant surface is a cone, and the g = constant surface is a plane. As in the other coordinate systems, the base vectors f, 6, and § are defined as unit vectors perpendicular to the constant r, 6, and ¢ surfaces, respec- tively, and pointing in the direction of increasing coordinate value. These base vectors, which are illustrated in Fig. 1.18, are mutually orthogonal, as they are in the other two coordinate systems. A right-handed spherical coordinate system is defined by t x 6 = 6. In the spherical coordinate system the differential lengths are dr, r d@, and r sin @dy. These are the differential lengths resulting when r, 6, and g, respectively, are increased by differential amounts. Since @ and ¢ are angles, the differential lengths are not simply d@ and dy. Fig- ure 1.19 illustrates the r dé and r sin 6 dg differential lengths. The total differential arc length resulting when all three coordinates are increased by differential amounts is (dl)? = (dr)? + (r d@)? + (r sin 6 dy)? z = z SP lo Sao i x ny ae dg— . x x Per sinoay rsing Fig. 1.19 The differential lengths r d@ and r sin 6 dg. sec. 1.5 mathematical introduction 15 Fig. 1.20 The differential volume in spherical coordinates. The differential volume is given by dV = drrd6rsin 0 dy =r? sin 6 dr do dp (1.13) as shown in Fig. 1.20. And, of course, a vector may be represented in the spherical coordinate system by the form A(7,8,¢) = Ax(r,0,0)t + Ao(r,6,v)6 + A,(r,0,0)6 which is similar to the representation in the other two coordinate systems. Note that the base vectors change their directions with @ and ¢- This variation must be kept in mind in integrating and differentiating the base vectors in spherical coordinates. Because of this difficulty, it is not desirable to use spherical coordinates in certain types of problems unless there are powerful symmetry arguments to the contrary. 1.5 ADDITION AND MULTIPLICATION OF VECTORS REPRESENTED IN THE THREE STANDARD COORDINATE SYSTEMS There is an obvious similarity in the relationships between the variables and base vectors in the three coordinate systems. Forexample, t x § = 2, # x § = 2,andt x 6 = §are similar relations. Also notice the similarity of the expressions A= A.(z,y,2)% + A,(z,y,2)9 + Ad(z,y,2)2 A= Ad(r,9,2)8 + Ag(r,y,2)6 + Adlr,o,2)2 A= Ax(r,6,0)8 + Aa(r,0,0)8 + Aol(r,0,0)6 A little reflection shows that there are similarities in all the expressions we have discussed. Let us take advantage of these similarities by defining three new generalized independent variables, w:, u2, and us, which stand for either 2, y, and z;r, y, and z; orr, 0, and y. In addition, let us define fy, de, 16 introduction to modern electromagnetics chapter one and Gs; as being three new generalized unit base vectors which stand for either %, §, and 2; f, $, and 2; orf, 6, and §. Then a vector represented in any of the three coordinate systems may be written in the generalized form A = Aj(us,ti2,ts) fr + Aa(us,t2,Us) fe + As(u1,t2,s) Os where A, stands for Az, A,, or A,,andsoon. Then we have the relations ti-ti=1 1-8-0 §f)-) <0 fi x fe = fs and A, =f)-A Since uz and us may be angles, additional definitions are required in order to represent differential lengths and differential volumes. Let us define dh = hy dur dl, = hy dus (1.14) dl; = he dus where dl, diz, and dl; are the differential lengths resulting when w, u2, and us, respectively, are increased by dui, due, and dus, respectively. The quantities h1, he, and h; are scale factors needed to make Eqs. (1.14) valid. There will be a different set of h’s for each coordinate system. From the discussion of differential lengths given previously, it is apparent, for exam- ple, that in cylindrical coordinates, dl, = dr and Ae d,=rdp and he=r dl, = dz and hy=1 Table 1.1 summarizes the scale factors for the three coordinate systems. Table 1.1 Summary of the scale factors Rectangular Cylindrical Spherical coordinates coordinates coordinates Ay h=1 ha = hp=r As=1 hy =rsin @ sec 15 mathematical introduction 17 The total differential arc length is given by (al)? = (dls)* + (dls)? + (dls)* = (hi dus)? + (he dus)? + (hs dus)? It follows easily that the differential volume is given by dV = hy duy he dus hs dus Vector addition may be conveniently discussed in terms of the repre- sentations in any of the three coordinate systems. Suppose we wish to add two vectors A and B with the general representations A= Aid + Ast, + Ast B = Bit + Bef: + Batts where, for brevity, the functional dependence of the vector components on Ux, U2, and uz has not been wri.ten explicitly. Then A+B = AiG, + Ard: + Acts + Bit, + Bot, + Bats Since vector addition is associative, this may be written as A+B = (Arf, + Bits) + (Ast: + Botte) + (Arts + Batt) From the graphical method of vector addition shown in Fig. 1.3, it follows that Arf + Bid = (4, + Bi), and therefore A+B = (Ai + Bi); + (Az + Bs)fy + (As + Bi) fs (1.15) It is apparent from Eq. (1.15) how the representation of vectors in coor- dinate systems simplifies addition. Next consider the dot product of A and B, A+B = (4,0, + As: + Ast) - (By, + Bod, + Bats) Since the dot product is distributive [see Eq. (1.6)], we write A+B = Asti: (Bid: + Bolt + Bas) + Asfy- (B10, + Bod, + Batts) + Asta (Bid, + Bofr + Batts) which can be expanded again to obtain terms such as A;f,- By0,. From Eq. (1.4) it can be shown that Ajf, + Byf; = A,B,0,- O, Because the unit vectors are mutually orthogonal, A+B reduces to the very important result A-B = A,B, + A2B, + A;B; (1.16) 18. introduction to modern electromagnetics chapter one The distributive property of the cross product may be used in a similar way to expand A x B, but care must be taken to keep the correct order, because the cross product is not commutative. Thus AXB = (Ait + Aste + Ass) x (Bid: + Bots + Bets) With the cross-product relations of the base vectors, A x B reduces to AXB = (A2B; — BzAs)&, + (AsB: — BsAi) 2 + (AiB2 — Bi Az) as (1.17) Equation (1.17) can be conveniently remembered as the expansion of the determinant, a & AXB=jA, A, As (1.18) |B, B, Bs; It should be noted that, in general, the relations between f,, fiz, and fi; are true only at a given coordinate point, because the base vectors can be variable vectors. For example, in cylindrical coordinates, f is a func- tion of y. It is obvious that f(y) -#(¢) =1 but £(¢1) + F(¢2) # 1 (1.19) as can clearly be seen from Fig. 1.21, This point isimportant with respect to the use of Eqs. (1.15), (1.16), and (1.18). For example, let us find the dot or cross product of the two vectors shown in Fig. 1.22 in a cylindrical coordinate system. We might be tempted to write A=Ai B= Bi x Fig. 1.21 Illustration of Fig. 1.22 Vectors used in the vector of Eq. (1.19). example of dot and cross products in a cylindrical coordinate system.

You might also like